09.04.2013 Views

2012 COURSE DATES: AUGUST 4 – 17, 2012 - Sirenian International

2012 COURSE DATES: AUGUST 4 – 17, 2012 - Sirenian International

2012 COURSE DATES: AUGUST 4 – 17, 2012 - Sirenian International

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

ECOLOGY, BEHAVIOR & CONSERVATION OF MANATEES & DOLPHINS AND THE<br />

COMPLEX MANGROVE-SEAGRASS-CORAL HABITATS WE SHARE WITH MARINE<br />

MEGAFAUNA WITHIN THE BELIZE BARRIER REEF LAGOON SYSTEM<br />

<strong>2012</strong> <strong>COURSE</strong> <strong>DATES</strong>: <strong>AUGUST</strong> 4 <strong>–</strong> <strong>17</strong>, <strong>2012</strong><br />

OVERVIEW & BACKGROUND<br />

Antillean manatees (Trichechus manatus manatus) are found throughout Central America and the Caribbean, but<br />

are Red Listed by the IUCN as endangered, in continuing decline, with severely fragmented populations. The UN<br />

Caribbean Environmental Programme considers them an endangered and protected species of regional concern,<br />

threatened by poaching, boat strikes, entanglement in fishing gear, and habitat degradation. Belize may be the last<br />

stronghold for Antillean manatees in the Caribbean; and the Drowned Cayes area is one of the most important<br />

activity centers in Belize. However, the existing body of knowledge is inadequate to develop and implement sitespecific<br />

management and recovery plans. Because manatees are elusive, endangered, and have slow reproductive<br />

rates, long-term studies in this area are necessary to evaluate and monitor the population status and develop practical<br />

conservation plans to ensure survival of the population, and ultimately, the sub-species. Our comprehensive and<br />

collaborative project began in October 1998 and hopefully, will continue indefinitely into the future!


Bottlenose dolphins (Tursiops truncatus) are not<br />

endangered, but their stocks are considered depleted<br />

by the U.S. Marine Mammal Protection Act. A<br />

study of the local population was started in the<br />

Drowned Cayes by Oceanic Society Expeditions<br />

(OSE) in 1997, however it was abandoned in 2001<br />

when OSE moved their base to Turneffe, an atoll<br />

approximately 10 miles east of the Belize Barrier<br />

Reef. We re-started a photo-id project in 2005 with<br />

the hope of additional collaboration with OSE and<br />

other dolphin researchers in the region.<br />

Although this project does not specifically<br />

investigate sea turtles in Belize, loggerhead,<br />

hawskbill, and green sea turtles are occasionally<br />

sighted during our expeditions. And, in 2011, our<br />

visiting scientist is a turtle expert!<br />

Students and volunteers have been an integral component of the project since its conception. We rely on you to<br />

assist us with a wide range of data collection and processing during each survey. During this course, half-day (3-4<br />

hours) manatee & dolphin surveys will be conducted from a small boat on 10 of your 14-day expedition.<br />

Additionally, each student will undertake an individual research project to investigate another component of our<br />

field site, exposing students the broader Conservation Biology issues (see syllabus for suggested research topics.<br />

WELCOME LETTER<br />

While in Belize, be warned, “If ya drink de watter, ya mus com bok.” This old Belizean Creole proverb is true! On<br />

my first trip to Belize in 1998, I drank the water and fell in love with both the people and the place…and the<br />

manatees, dolphins, and other wildlife of<br />

course! Since that first trip, Belize has been<br />

my 2 nd home. Hopefully, you will soon be<br />

equally enthusiastic about the opportunities and<br />

challenges that await you in this tropical<br />

paradise.<br />

This is no ordinary field course! During the<br />

course, you will be collecting data on<br />

manatees, dolphins, and their habitat that my<br />

research partners and I actually use for our<br />

long-term studies of Antillean manatees and<br />

bottlenose dolphins in the Drowned Cayes<br />

(pronounced "keys") area of Belize. Although<br />

our project focuses primarily on manatees and<br />

dolphins, you will also experience the<br />

mangrove-seagrass-coral reef system in which<br />

we live and work. Perhaps most importantly,<br />

the data are also used by Belizeans, local<br />

NGOs, agencies, and decision-makers, to<br />

better develop marine conservation strategies.<br />

The Drowned Cayes is a complex maze of<br />

mangrove islands, just east of Belize City,<br />

surrounded by the Caribbean Sea. Since we<br />

are just two miles inside the Belize Barrier<br />

Reef, we are quite well protected from severe<br />

wave action, but if a tropical storm or hurricane threatens the western Caribbean, we will implement our Hurricane<br />

Plan, which includes evacuation. The Caribbean Sea and sunshine are breathtaking, but they can cause great<br />

Belize Field Course Briefing Page 1 4/4/<strong>2012</strong>


discomfort without proper protection, so you must pay close attention to the Packing Checklist in this briefing.<br />

Photo above: (c) C. Self-Sullivan: “Mario” a friendly, unmarked manatee videoed at North Gallows in 2001.<br />

You should be prepared to spend long periods of time outdoors, on the island and on the water in a small boat,<br />

searching for and observing manatees and dolphins. If you wish, you will have an opportunity to snorkel in the<br />

seagrass beds, mangrove bogues, or on coral patches. But please don’t expect to swim with manatees and dolphins <strong>–</strong><br />

our wild animals are not acclimated to humans and intentional swimming with manatees & dolphins is not permitted<br />

in Belize. After a few days in the field, you will understand why we refer to manatees as elusive. We can pretty<br />

much guarantee that you will see both manatees & dolphins in the wild, but we can’t predict how many or how close<br />

they will come to our research boat. However, the more you read about manatee & dolphin ecology and behavior<br />

prior to your arrival, the more rewarding your observations will be.<br />

You should also be prepared for total immersion into private island living. As field researchers, we spend much of<br />

our time innovating and problem solving. On this<br />

expedition, you will experience the unique lifestyle<br />

of wildlife researchers under moderate conditions.<br />

While we live comfortably compared to most cayedwellers,<br />

it’s quite different from life in the northern<br />

hemisphere. Spanish Lookout Caye, a 184-acre<br />

mangrove island, is shared by Spanish Bay<br />

Conservation and Research Center, the Hugh Parkey<br />

Foundation for Marine Awareness and Education,<br />

and Hugh Parkey’s Belize Dive Connection and<br />

Adventure Lodge. During your two recreational<br />

days, we will be using Belize Dive Connections to<br />

experience an inland adventure to a Maya site, and<br />

a sea adventure on kayaks, snorkel or SCUBA (you<br />

may SCUBA only if you are a certified diver...bring<br />

your sea card & dive log). Photo (c) C. Self-<br />

Sullivan: “Dubya” a dolphin we’ve observed in our<br />

study area since 2004.<br />

Evening entertainment includes good conversation,<br />

star gazing, and a few old fashioned board and card<br />

games. So bring your favorite stories to share with us as we create some new stories to take back to your friends!<br />

Feel free to bring non-electronic musical instruments and games, too. Hopefully, participating on this expedition<br />

will leave you with a new perspective on sustainable living <strong>–</strong> are you up to the challenge?<br />

Cheers,<br />

Caryn Self-Sullivan, Ph.D.<br />

Adjunct Faculty, Nova Southeastern University<br />

President & Co-founder, <strong>Sirenian</strong> <strong>International</strong><br />

200 Stonewall Drive, Fredericksburg, VA 22401-2110<br />

Email: cselfsullivan@sirenian.org<br />

Voice: 540.287.8207 | Fax: 888.371.4998<br />

Belize Field Course Briefing Page 2 4/4/<strong>2012</strong>


TABLE OF CONTENTS PAGE<br />

BACKGROUND & PROJECT OVERVIEW Cover - 1<br />

WELCOME LETTER 1-2<br />

REGISTRATION FORM APPENDIX<br />

<strong>COURSE</strong> SYLLABUS (REQUIRES SIGNATURE) APPENDIX<br />

<strong>COURSE</strong> POLICY AND LIABILIT RELEASE FORM (REQUIRES SIGNATURE) APPENDIX<br />

REGISTRATION CHECK LIST 3 (BELOW)<br />

ABOUT BELIZE 4<br />

PRINCIPAL INVESTIGATOR, CO-PI, VISITING SCIENTISTS 4-5<br />

FIELD TRAINING AND ASSIGNMENTS 6-7<br />

ACCOMMOATIONS & FOOD 7-8<br />

ENVIRONMENTAL CONDITIONS 8<br />

POTENTIAL HAZARDS 9<br />

MEDICAL CONDITIONS OF SPECIAL CONCERN 9<br />

HEALTH INFORMATION 10<br />

PACKING CONSIDERATIONS 10<br />

EMERGENCIES IN THE FIELD 10-12<br />

OTHER USEFUL INFORMATION 12<br />

PACKING CHECK LIST 13-14<br />

APPENDIX (Forms, Data Sheets, Background Readings, Course Flyer, CVs) 15<br />

REGISTRATION & Pre-FIELDING CHECK LIST Session I<br />

Sessions II & III 1 st Group of items is due 90 days prior to fielding<br />

2 nd group of items is due 60 days prior to fielding Sent Date<br />

1 May <strong>2012</strong> Completed Registration Form _________<br />

1 May <strong>2012</strong> Deposit Invoice Paid _________<br />

1 May <strong>2012</strong> Signed Course Policy & Liability Release Form _________<br />

1 May <strong>2012</strong> Signed Syllabus & Unofficial Copy of Transcripts _________<br />

1 May <strong>2012</strong> Low-income Country Scholarship Essay _________<br />

1 May <strong>2012</strong> 1 st, 2 nd, and 3 rd Choice of Belize: Sea to Stars Topic _________<br />

1 May <strong>2012</strong> 1 st, 2 nd, and 3 rd Choice of Independent Research Topic _________<br />

1 June <strong>2012</strong> Balance Due Invoice Paid _________<br />

1 June <strong>2012</strong> Copy of Passport & Current Photo (Head Shot) _________<br />

1 June <strong>2012</strong> Copy of DAN Preferred Membership Card _________<br />

1 June <strong>2012</strong> Copy of SCUBA Certification Card (if diving) _________<br />

1 June <strong>2012</strong> Copy of Airline Travel Itinerary & Receipt _________<br />

1 June <strong>2012</strong> Proposed Presentation Format for Belize: Sea to Stars _________<br />

1 June <strong>2012</strong> Literature Review & Citations for Independent Research Topic _________<br />

Belize Field Course Briefing Page 3 4/4/<strong>2012</strong>


ABOUT BELIZE<br />

Research Site: The project is based on Spanish Lookout Caye, a 184-acre twin-mangrove island located in the<br />

Drowned Cayes, about 10 miles east-southeast of Belize City. The Drowned Cayes are a maze of mangrove islands<br />

inside the Belize Barrier Reef. Approximately 10 acres of Spanish Lookout Caye have been filled and developed,<br />

leaving over <strong>17</strong>0-acres of pristine mangroves and<br />

mangrove swamps divided by a fast flowing channel<br />

of water commonly known as Gilroy’s Creek.<br />

Mangrove islands are famous for their wildlife and<br />

you will share this island home with birds, crabs,<br />

mosquitoes, and sandflies; you might even be lucky<br />

enough to sight an American crocodile or boa<br />

constrictor. Thirty to forty hours will be spent on our<br />

research boat, and marine hazards in the area include<br />

jellyfish and fire coral. The tropical sun is strong and<br />

the humidity is very high. Among the best reasons for<br />

visiting Belize are its long-standing and rapidly<br />

expanding conservation ethic and the incredible<br />

diversity of natural habitats. The spectacular Belize<br />

Barrier Reef, which is the second largest barrier reef in the world and runs along the entire 280 km coastline<br />

(Beletsky 1999), is perhaps its most celebrated natural treasure.<br />

Cultural, Social and Political Environment: Belize is a small country located between Mexico and Guatemala on<br />

the Yucatan Peninsula and was formally known as British Honduras. Because Belize was a British protectorate,<br />

English is taught and spoken at all school levels and is the official language. The more common language spoken<br />

between Belizeans is Creole <strong>–</strong> an unwritten language that combines English with African, Maya, and Spanish words.<br />

Most Belizeans speak at least two languages: English and at least one other such as Creole, Garifuna, Spanish, Maya<br />

Mopan, Kechi Maya, German (Menonites), etc.<br />

Belize is a politically stable democracy, with a Parliamentary system of government and elections every five years.<br />

Elections were held in spring of 2008, and the United Democratic Party (UDP) currently holds the majority of<br />

governmental seats. The People’s United Party (PUP) is the dominate opposition, having just lost their first election<br />

in 10 years. There are a few independents in office. The current Prime Minister is the Honorable Dean Barrow (also<br />

father of Rapper Jamal “Shyne” Barrow). Belizeans are quite vocal about their political opinions; almost everyone<br />

turns out to vote, and local taxi drivers are a good source of information about the current issues. Christianity is the<br />

predominant religion with many churches serving both Catholic and Protestant congregations.<br />

Like any large urban area, Belize City has its share of crime. The relatively small population of less than 300,000<br />

people consists of predominantly five unique cultures: Maya, Mistiso, Creole, Garifuna, and European. More<br />

recently Mennonites, Chinese, and Taiwanese populations are growing in Belize.<br />

PRINCIPAL INVESTIGATOR<br />

Caryn Self-Sullivan, Ph.D., is currently adjunct faculty at Nova<br />

Southeastern University, President of <strong>Sirenian</strong> <strong>International</strong>,<br />

Scientific Advisor to the NCRC West African manatee project in<br />

Volta Lake, Ghana, and Marine Science Advisor to the Hugh<br />

Parkey Foundation for Marine Awareness and Education in Belize.<br />

She graduated from Coastal Carolina University in 1997 with a<br />

B.S. in Marine Science and minors in Mathematics and Biology.<br />

She was an NSF Graduate Fellow from 1999-2001, and received<br />

her Ph.D. in Wildlife & Fisheries Sciences in 2008. Additionally,<br />

Dr. C serves on the IUCN Species Survival Commission Sirenia<br />

Specialist Group, the Belize National Manatee Working Group, and<br />

the Belize Marine Mammal Stranding Network. She has taught<br />

university level courses in Environmental Education, General<br />

Biology and Conservation Biology. She has taught professional<br />

level workshops focused on the order Sirenia in the Dominican<br />

Dr. Self-Sullivan<br />

Belize Field Course Briefing Page 4 4/4/<strong>2012</strong>


Republic, Belize, Ghana, and the USA. Her research interests include marine biology, animal behavior, endangered<br />

marine species, and conservation biology, with a focus on marine mammals. In the field, Dr. C is responsible for<br />

supervising the students and staff, field training, coordinating the lecture/learning/discussion series, overseeing<br />

experimental design and data collection methods, capturing manatees with underwater video camera equipment, and<br />

capturing dolphins with above water digital camera equipment. During this course, Dr. C will be your primary<br />

expert on the local research project, manatees, and dolphins.<br />

CO-PI AND VISITING SCIENTISTS (CVS ARE IN THE APPENDIX SECTION)<br />

Katherine S. LaCommare, Ph.D., Co-PI will receive her<br />

PhD from the Environmental Biology Program, Department of<br />

Biology, University of Massachusetts, Boston, in May 2011!<br />

She is adjunct faculty at Lansing Community College in<br />

Michigan. Her previous degrees include a M.S. in Forestry,<br />

Conservation Biology Program, Department of Forestry and<br />

Natural Resources, Purdue University; and a B.S. in<br />

Anthropology/Zoology, University of Michigan. Katie cofounded<br />

<strong>Sirenian</strong> <strong>International</strong> with Dr. C and has been the co-<br />

PI on our long-term manatee research project since its<br />

inception in 1998. Unfortunately, Katie may not be able to<br />

join us during the 2001 field course.<br />

Heather J. Kalb, Ph.D., Visiting Scientist is an Assistant<br />

Dr. LaCommare<br />

Profession of Biology at West Liberty University in Wheeling,<br />

WV. She received her Ph.D. in Zoology from Texas A&M<br />

University in 1999. Dr. Kalb specializes in the reproductive<br />

physiology of turtles. She has taught university level courses in Zoology, Animal Diversity, General<br />

Biology, Scientific Communication, Comparative Vertebrate Anatomy & Physiology, Animal Behavior,<br />

Vertebrate Zoology, Conservation Biology, and Vertebrae Ecology. During this course, Dr. Kalb will be<br />

your primary expert on sea turtles and the fields of zoology, vertebrate anatomy, physiology.<br />

Bruce A. Schulte, Ph.D. Visiting Scientist, is the Biology Department Head at Western Kentucky<br />

University. He received his Ph.D. from the State University of New York, College of Environmental<br />

Science and Forestry in 1993; his M.S. in Biology from University of Southern California, and his B.S.<br />

from the College of William and Mary. He has regularly taught courses in Animal Behavior, Behavioral<br />

Ecology, Chemical Ecology, Conservation, and Environmental Biology. His research interests include<br />

communication and social behavior of herbivorous mammals, such as elephants, manatees, beavers and<br />

horses. His research group also examines how an understanding of behavior can facilitate positive<br />

human-animal interactions, such as reducing human-wildlife conflict. When in the field, Dr. Schulte will<br />

be your primary expert in the fields of animal behavior, chemical ecology, and conservation behavior.<br />

Jessica R. Young, Ph.D., Visiting Scientist is Associate Professor of Biology and Associate Vice<br />

President for Academic Affairs at Western State College of Colorado. She received her Ph.D. in<br />

Population Biology and<br />

Behavioral Ecology from Purdue University in 1994; and her B.A. in Ecology, Behavior, and Evolution<br />

from UC San Diego in 1988. Her general research interests integrate evolutionary theories of behavioral<br />

ecology and animal communication with applied aspects of conservation biology and wildlife<br />

management. For the past two decades, she has been working with a unique species of grouse, the<br />

Gunnison Sage-grouse, which was recognized by the AOU in 2000 as a distinct species based on<br />

physical, behavioral, and genetic traits and the first new bird species described in over 100 years. In the<br />

field, Dr. Young will be your primary expert in the fields of behavioral ecology and ornithology.<br />

Belize Field Course Briefing Page 5 4/4/<strong>2012</strong>


FIELD TRAINING AND ASSIGNMENTS<br />

During each field course, students are encouraged to<br />

become proficient in at least one aspect of the long-term<br />

research project and will have input on the assignments<br />

in which they would like to participate. For example,<br />

one participant might become proficient recording<br />

behavioral data, while another might be better at<br />

operating the Global Positioning System (GPS) or<br />

recording environmental data. Most days, students will<br />

spend 1/2 day (3-4 hours) on a small boat conducting<br />

surveys and observing manatee & dolphin behavior.<br />

Our data collection is generally done from 25-foot<br />

fiberglass research boats, equipped with a single 115 HP<br />

Yamaha four-stroke engine. We carry a mobile phone, a<br />

magnetic compass, GPS unit, fire extinguisher, First Aid<br />

kit, and life jackets for 13 passengers, including the<br />

captain. Other standard safety equipment includes an<br />

anchor with anchor line, swim ladder, pole, bailing<br />

bucket, and bimini top for shade. There is no head<br />

(toilet facility) on board. We relieve ourselves in the<br />

water or in a bucket.<br />

Skills and talents extremely helpful to the project<br />

include patience, flexibility, attentiveness, a love for<br />

watching animals, a passion for learning, a passion for<br />

living sustainably in an outdoor environment,<br />

swimming, snorkeling, good handwriting, positive<br />

problem solving, positive attitude, strong team spirit, respect for low-tech data collection methods, and the ability to<br />

design and create something from nothing.<br />

Students will be expected to participate in most of the following tasks:<br />

• Continuous scanning surveys: These include both boat surveys and point scans designed to search for manatees<br />

& dolphins within the study area. During boat surveys and point scans students are needed to actively watch for<br />

manatees, dolphins and signs of their presence, and to record survey and sighting data. During sightings,<br />

students will record additional data, including behavioral states, breath cycles, and movements, as well as data<br />

about other boats in the scan area.<br />

• Environmental variable recording: Air and water temperature, salinity, wind direction and speed, sea state,<br />

turbidity, water depth, and bottom type will all be recorded. During boat surveys, point scans, and/or whenever<br />

a manatee or dolphin is sighted, locations will be marked with a GPS unit and general location will be plotted<br />

on the field map. At the end of each point scan students will use field equipment to take these measurements.<br />

• Focal follows: During focal follows, which are 40-minute increments of time during which we focus on a single<br />

manatee or group of dolphins, students record manatee or dolphin behavioral states, breath cycles, movements,<br />

and any other activity in the area.<br />

• Underwater ID captures: These are attempted by the Principal Investigator (PI) only, using a digital video<br />

camera and underwater housing. During an underwater attempt, students record both the manatee and the PI<br />

movements on a focal follow data sheet.<br />

• Surface photo-identification: During dolphin focal follows, the PI photographs dolphins from the bow of the<br />

research vessel; students document details on data sheets. Students will have an opportunity to photo dolphins<br />

when conditions are appropriate.<br />

Note: Students should NOT expect to swim with the manatees and dolphins, as commercial swimming with these<br />

endangered animals is prohibited in Belize. The PI has special permission to get into the water to photograph and<br />

video-tape animals, but this permission does not extend to students.<br />

Belize Field Course Briefing Page 6 4/4/<strong>2012</strong>


ACCOMMODATIONS<br />

At Spanish Bay, students and staff will share dormitory style housing. Each room will consist of a combination of<br />

bunk beds, sleeping up to 12 per room. Shared bathroom facilities include conventional toilets, showers and sinks.<br />

Room assignments will be determined by gender. Private rooms for couples are not available. Rooms have<br />

screened windows to allow for ocean breezes and limit nocturnal insect visitors; however we recommend you<br />

consider bringing your own personal mosquito netting, as well as duct tape and string for hanging net over your<br />

bunk.<br />

Spanish Lookout Caye is powered by solar, wind, and diesel<br />

generators. Fresh water is precious and we mandate water<br />

conservation. We capture rainwater and recycle all waste water<br />

using a state-of-the-art tertiary recycling plant. Volunteers are<br />

taught to conserve water by taking one quick (island) shower<br />

per day, turning off the faucet while lathering soap, brushing<br />

teeth, etc. Drinking water is purchased from Belize and<br />

available 24-7 at designated dispensers.<br />

There is sufficient power to recharge batteries. We encourage<br />

students to bring digital cameras. We rely on you to take<br />

pictures of the project and we like to download them during the<br />

expedition. Please bring the appropriate charging devices (110<br />

volts AC, 60 Hz, flat two-pin plugs) and your data cable to<br />

share images with your classmates and instructors. Also, please<br />

bring rechargeable batteries; used disposable batteries must<br />

return home with you! Note that there is NOT adequate electricity for high voltage appliances such as curling irons,<br />

blow dryers, or coffee makers. Internet access and phone usage on the caye is limited to emergency use by the<br />

instructors. Many US and UK cell phones (except Sprint) work in Belize, but the international fees are generally<br />

expensive. SMS Texting works well and costs about $0.50 per message.<br />

NOTE: In <strong>2012</strong>, we will also spend a few days on Corozal Bay in northern Belize with the<br />

Sarteneja Alliance for Conservation & Development Home Stay Program and may visit other<br />

conservation sites.<br />

FOOD<br />

At Spanish Bay, we will eat in the Belize Adventure Lodge dining area. Due to the logistics of living on a<br />

mangrove island, there is a limited amount of food variety during the expedition. Some special diets may be<br />

accommodated if advanced notice is given. Generally we eat a lot of chicken! BE SURE TO RECORD ANY<br />

DIETARY RESTRICTIONS ON YOUR REGISTRATION FORM…INCLUDING, BUT NOT LIMITED TO<br />

VEGETARIAN, VEGAN, LACTOSE INTOLERANT, GLUTEN INTOLERANT, FOOD OR OTHER<br />

ALLERGIES. The goal is to make our diet as traditional as possible, in order to make food a part of your<br />

experience! Weekly menus may include eggs and chicken, beef, pork, and local seafood (rarely), seasonal fruits and<br />

vegetables, tortillas, rice and beans, beans and rice (there is a difference), and pasta dishes. Water is the staple drink,<br />

supplemented with fruit juices, coffee, and tea. Also, if there are foods you do not eat for any reason please let us<br />

know so we don’t waste food by putting items on your plate that you won’t eat. For example, tomatoes are<br />

expensive so if you don’t eat them, just let us know and the staff will not put them on your plate!<br />

Comfort foods/drinks such as chips, cookies, candy, sodas, beer, rum, and wine are available (at your own expense)<br />

on the caye. A Gift Shop with limited supplies is located on our caye and is generally open daily from 9-4. Below<br />

are examples of the foods you might expect during the expedition. Please bear in mind that variety depends on<br />

availability.<br />

Breakfast: Eggs, beans, breakfast meats, tortillas, pancakes, oatmeal, fresh fruits<br />

Lunch: Tortillas, pasta, fresh veggies, rice, beans, chicken, boiled eggs, fresh fruit<br />

Dinner: Rice, beans, chicken, beef, pasta, veggies<br />

Snacks: Oranges, bananas, papaya, pineapple, watermelon<br />

Beverages: Water, coffee, tea, fruit juices<br />

Belize Field Course Briefing Page 7 4/4/<strong>2012</strong>


ENVIRONMENTAL CONDITIONS<br />

SBCRC is situated on a 184-acre twin mangrove island, partially cleared and filled, but with over<br />

80% of the mangrove ecosystem intact, including the mosquitoes and sand flies. A mangrove<br />

island harbors many little hazards such as broken shells to cut your feet, and mangrove roots to<br />

trip you up, so it’s important to pay close attention to this rustic environment, which is free of<br />

paved roads and sidewalks.<br />

The sun is very strong here, and<br />

brief periods of intense rain are not<br />

uncommon during the field season.<br />

More extreme tropical storms and<br />

hurricanes traditionally occur from<br />

June through November with late-<br />

August, September, and October as<br />

the most active periods.<br />

In the event of a hurricane, we may<br />

evacuate the island and move inland<br />

for the duration of the storm. This<br />

has occurred three times during the<br />

14 years of this project.<br />

Belize Field Course Briefing Page 8 4/4/<strong>2012</strong>


Potential Hazards<br />

We take pride in our experience, training, and track record with respect to students’ health and safety. So even when<br />

it seems that we are “mothering” you too much, we do expect you to follow our advice regarding your healthy and<br />

safety at all times!<br />

Hazard Type Associated Risks and Precautions<br />

Climate Students must be prepared to spend long hours in hot, humid, and wet conditions. The<br />

tropical sun is very strong. Dehydration, sunburn, and other heat related illnesses are a risk.<br />

Insects Sand flies and mosquitoes can be problematic. Sand flies are believed to be a vector for<br />

leishmaniasis in some regions. Mosquitoes may transmit a number of diseases (see<br />

Diseases below). Bot flies are also found in Belize, and mosquitoes may transmit their<br />

larvae to human hosts where the larvae will grow and develop. This is not life threatening,<br />

but can be painful and unpleasant.<br />

Marine life Fire corals, several species of stinging jellyfish, sharks, sea urchins, lionfish, and other<br />

potentially dangerous marine organisms can be found in the study area. These can all give<br />

painful and occasional severe stings or bites, which may become infected if untreated.<br />

Those with a dangerous allergy to bee or wasp stings may have a similarly dangerous<br />

reaction to corals and jellyfish. The best prevention is to avoid and not touch the animals.<br />

Please bring your epi-pen if you are allergic to anything.<br />

Snorkeling All the inherent risks of snorkeling are obviously present, including the effects of<br />

environmental conditions, marine life and other risks specific to your own physical/medical<br />

history. Snorkeling is optional and will be conducted in seagrass beds and on coral<br />

patches. Volunteers who chose to snorkel should know how to do so safely without<br />

hyperventilating or kicking up the substrate.<br />

Boats We will be aboard a boat for most fieldwork. All the fiberglass boats should have ladders<br />

and a bimini cover for sun protection. However, in some instances we may use a boat that<br />

lacks a ladder or shade. Deck surfaces of boats will become slippery and may place you at<br />

risk of slips, falls, and injuries that result from these accidents.<br />

Disease Diseases found in tropical regions include malaria, dengue fever, filariasis, leishmaniasis,<br />

onchocerciasis, trypanosomiasis (Chaga’s disease), schistosomiasis, leptospirosis, rabies,<br />

brucellosis, hepatitis, and typhoid. Most diseases are prevented with basic safety cautions.<br />

Driving Driving conditions are considered poor by western standards and pose inherent risks.<br />

Students will not be permitted to drive during the expedition.<br />

Medical Conditions of Special Concern<br />

Students should be physically fit, competent swimmers, and comfortable spending 3-4 hours on a boat. Those with<br />

chronic back problems or seasickness will find working and riding in small boats very uncomfortable. If you suffer<br />

from seasickness and intend to treat this with either over-the-counter or prescribed medication, please discuss the<br />

use and side effects with your physician and notify the PI before fielding. Be sure to tell your doctor that you will<br />

be spending all day in the sun! Please also let the PI know what medications, if any, you are taking for seasickness<br />

and/or malaria prevention. Some prophylactics have severe interactions with sun exposure and must be avoided.<br />

Any conditions that interfere with or limit stamina in the water, balance, swimming, or breathing should be carefully<br />

considered. If you have a current ear or sinus infection, it should be fully healed prior to participation in snorkeling<br />

or SCUBA. If you are allergic to bee stings, you may be allergic to Cnidaria stings (jellyfish and corals) and must<br />

bring an Epi-kit. Visual acuity (corrected via glasses or contacts is fine) and good hearing are important.<br />

Conditions or medications that increase one’s light sensitivity or sunburn risk should be discussed with a physician.<br />

Belize Field Course Briefing Page 9 4/4/<strong>2012</strong>


HEALTH INFORMATION<br />

Routine Immunizations<br />

All volunteers should make sure to have the following up-to-date immunizations: DPT (diphtheria, pertussis,<br />

tetanus), polio, MMR (measles, mumps, rubella) and varicella (if you have not already had chicken pox). Please be<br />

sure your tetanus shot is current.<br />

Project Inoculations & Prophylactics<br />

The following are recommendations only. Medical decisions are the responsibility of each student and their doctor.<br />

Note that health conditions around the world are constantly changing, so keep informed and consult your physician,<br />

a local travel health clinic, the US Center for Disease Control (www.cdc.gov), the World Health Organization<br />

(www.who.int). Please consult your physician for guidance on inoculations if you intend to travel to other parts of<br />

the country.<br />

Typhoid<br />

Hepatitis A<br />

Hepatitis B<br />

These inoculations are recommended by the CDC for health reasons whenever you are<br />

traveling outside the USA.<br />

Malaria A malaria prophylaxis that allows exposure to the sun is recommended by the CDC<br />

Other Advice / Information<br />

• Malaria: Malaria is not present at the research site, but it is found within Belize. A prophylaxis is<br />

recommended by the CDC for all areas except Belize City. The risk is highest in the western and southern<br />

regions of the country, which you may visit on a recreational day.<br />

PACKING CONSIDERATIONS<br />

Remember to review the Packing Checklist at the end of this briefing.<br />

General Considerations<br />

Do not bring more luggage than you can carry and handle on your own. Space is also extremely limited in the dorm<br />

rooms. We recommend that you pack a carry-on bag with an extra set of field clothing (shorts, t-shirt, hat, swim<br />

suit, mask and snorkel) and personal essentials in the event that your luggage is lost and/or takes several days to<br />

catch up with you.<br />

Remember to bring old t-shirts and shorts that you don’t mind getting dirty and possibly ruining. Expect to wear the<br />

same shorts and t-shirts repeatedly due to lack of laundry facilities. Clothes get ruined in the field; you will NOT<br />

need any good/nice clothes at the research camp. You might want to save a clean t-shirt and shorts for visiting<br />

inland sites, but, you don’t need dress slacks or skirts. If you have side trips planned before/after your research trip,<br />

plan accordingly.<br />

We have found some diversity among our previous students as to what they think should be mandatory. It varies not<br />

only with temporal conditions on the island, but also with student comfort when traveling within the developing<br />

world. The Packing Checklist at the end of this briefing contains recommendations based on our experience with<br />

over 500 previous students & volunteers. Some folks will require more creature comforts, and others can do without<br />

some of the recommended items. All volunteers should read this entire briefing carefully, and those who have done<br />

a lot of traveling can use the information to pack according to their experience. Less experienced travelers should<br />

bring everything we recommend!<br />

Note: As the project is stationed on a mangrove island, any trash produced must be burned. There are few recycling<br />

facilities in Belize. Please help protect the environment by leaving disposable products and plastic packaging at<br />

home.<br />

Belize Field Course Briefing Page 10 4/4/<strong>2012</strong>


Cultural Considerations<br />

Belize is predominately a Christian culture. Shorts and t-shirts are fine for both men and women. Swimwear is<br />

appropriate for beaches, but not for Belize City, where shirts and shoes are recommended at all times.<br />

Essential Items<br />

While it is recommended that you pack as light as possible, the following items are essential for participation: bug<br />

repellent for mosquitoes, oil (Avon Skin-so-Soft , baby oil, or olive oil is recommended) for sand flies, sunscreen<br />

(15-45 SPF), hat, long-sleeved shirt/cover up for boat, swimsuit, 1-2mm wetsuit or dive skin, and field clothes.<br />

Please see the Packing Checklist for a complete list of what you will need to take with you. We recommend<br />

going through the list with a pen or pencil and marking off each required item right before you leave for your<br />

expedition.<br />

EMERGENCIES IN THE FIELD<br />

The researchers and their host, Hugh Parkey’s Belize Dive Connection, are concerned with your health and safety<br />

during the field course. Minor injuries will be treated onsite using Red Cross First Aid and DAN Marine First Aid<br />

procedures. We take a precautionary approach to minor illnesses and injuries and will insist on scheduling an<br />

appointment with a local doctor if the situation does not improve within 24-48 hours after First Aid treatment.<br />

Volunteers with major illnesses or injuries will be transported to Belize City for medical advice and treatment.<br />

There is always a boat available for transport in case of an emergency. In case of a life-threatening illness/injury,<br />

we will take the following steps:<br />

1) Insure that all students/staff are safe from further injury<br />

2) Give First Aid/CPR as necessary to stabilize the victim(s)<br />

3) Contact HP's Belize Dive Connection to report the situation<br />

4) Transport the victim(s) by boat to Belize City or Contact DAN if indicated<br />

5) Notify victim’s Emergency Contact at first opportunity<br />

6) Arrange transportation from the dock in Belize City directly to an emergency care facility, Belize Medical<br />

Associates<br />

7) Follow up with HP's Belize Dive Connection as soon as the victim is under professional medical care<br />

8) Follow up with the rest of the class to keep them informed<br />

9) Follow-up with victim’s Emergency Contact<br />

OTHER USEFUL INFORMATION<br />

• Our Host: Ms. Teresa Parkey, Hugh Parkey's Belize Adventure Lodge, PO Box 1818, Belize City, Belize,<br />

Central America. Tel: ++501-223-4526 or ++501.223-5086 Fax: ++501-610-5235, E-mail:<br />

hugh@belizediving.com<br />

• Do not book your flight until your registration and the course has been confirmed! Expected confirmation date:<br />

60 days prior to fielding.<br />

• Airport Code BZE: American Airlines and Continental Airlines fly into Philip S. W. Goldson <strong>International</strong><br />

Airport in Belize City daily; Delta and US Airways also have a more limited schedule. If you are flying from<br />

the West Coast of the US, you might consider TACA. Round-trip airfare is currently running between US$600-<br />

$800.<br />

• You must have a Passport for entry into Belize; if you are a US Citizen you will automatically be granted a 30day<br />

tourist VISA upon arrival. If you are not a US Citizen, or for additional information, visit the Belize<br />

Tourism Board Website for additional information: http://www.travelbelize.org/<br />

• Consult the US State Department Website for current information regarding travel to Belize:<br />

http://www.state.gov/p/wha/ci/bh/index.htm<br />

Belize Field Course Briefing Page 11 4/4/<strong>2012</strong>


• Consult the CDC for immunization recommendations: http://wwwnc.cdc.gov/travel/destinations/belize.aspx<br />

• If you plan to arrive before the rendezvous date or remain in Belize after the departure date, please contact Dr.<br />

Self-Sullivan, for additional information on travel and accommodations in Belize. You may find the Moon<br />

Handbook Belize 8 th Edition, by my friend and colleague Josh Berman, a valuable resource, available online at<br />

Amazon.com.<br />

• Recommended Field and Travel Guides:<br />

Birds of Belize (The Corrie Herring Hooks Series), By H. Lee Jones<br />

Moon Handbooks Belize, 8 th Edition, by Joshua Berman<br />

Travellers' Wildlife Guides Belize & Northern Guatemala, By Les Beletsky<br />

Reef Creature Identification: Florida, Caribbean, Bahamas by Paul Humann and Ned DeLoach<br />

Reef Coral Identification: Florida, Caribbean, Bahamas (Reef Set, Vol. 3) by Paul Humann and Ned<br />

DeLoach<br />

Reef Fish Behavior: Florida, Caribbean, Bahamas by Paul Humann and Ned DeLoach<br />

Reef Fish Identification: Florida, Caribbean, Bahamas by Paul Humann and Ned DeLoach<br />

• BELIZE time zone: GMT/UTC -6:00; Daylight Savings Time is NOT observed in Belize.<br />

• Local currency: Belize dollars, however, US dollars are accepted everywhere for a fixed exchange rate of<br />

US$1 = BZ$2. There is NO NEED to change US$ into BZ$.<br />

• Electricity: 110 volts AC, 60 Hz, flat two-pin plugs (same as USA). Electricity is only minimally available at<br />

the project site, although charging of laptops, digital cameras and some small electronics is fine. Students DO<br />

NOT have Internet access during the course!<br />

• Language: English, also spoken: Spanish, Maya, Garifuna, Creole<br />

• Telephone dialing codes: When calling Belize from another country, dial the country’s international dialing<br />

code (e.g., 011 in the USA), followed by 501 (Belize Country Code) and the number (e.g., 223.4526). When<br />

calling within Belize, omit the 501, and just dial the number. When calling another country from Belize, dial<br />

00, followed by the other country’s country code and the number. For example, to call the USA, dial 001+area<br />

code+number. I will rent a local cell phone upon arrival and text you the number so you can contact me during<br />

your travel is necessary!<br />

• Personal conduct: The field course and project are able to remain in Belize due to the courtesy of the people<br />

and government of Belize. As such, you will be expected to conduct yourselves in a manner that is respectful of<br />

local sensitivities, customs, and laws. Any violations of Belizean law will be prosecuted in Belize with no<br />

recourse to foreign laws and attorneys. Any conduct that reflects negatively on project will be grounds for<br />

immediate deportation at the expense of those involved.<br />

• Personal funds: Past students have spent a wide range of funds during the course. You will have opportunities<br />

to shop on at least one of your recreational days. The airport also offers many gift shops in the departure area.<br />

Small bills are useful (US$1, US$5, US$10) as change will be given in BZ dollars. Traveler’s checks and credit<br />

cards are more difficult to use, but OK at more and more locations each year. Debit cards may be problematic,<br />

but VISA and MasterCard credit cards have been used to get cash advances from banks in Belize City. There<br />

are ATMs, but they appear tied to VISA credit cards only. Larger stores, restaurants and hotels take VISA and<br />

MasterCard credit cards, but few take American Express. Smaller shops and street vendors only take cash.<br />

• Tips: As a visitor, it is customary to tip for services in Belize. You are responsible for tipping anyone you think<br />

gave exceptional service, including but not limited to the field assistants, housekeepers, restaurant staff,<br />

bartenders, dive master, drivers, tour guides, and boat captains during recreational activities. We generally<br />

“pass the basket” for tips at the end of each team.<br />

Belize Field Course Briefing Page 12 4/4/<strong>2012</strong>


Essential Items<br />

PACKING CHECKLIST<br />

Photocopies of your passport, flight itinerary and credit cards in case the originals are lost or stolen; the<br />

copies should be packed separately from the original documents<br />

Passport and Driver’s License and other Photo ID such as your Student ID<br />

Required Items Necessary for Successful Completion of Coursework<br />

Laptop Computer with USB port (for pen drive)<br />

Digital Camera and Underwater Housing<br />

Copies of articles from your Literature Review<br />

Copies of all journal articles (included in this briefing)<br />

Clothing/Footwear for Fieldwork<br />

Old shorts & old t-shirts<br />

Lightweight, breathable long<strong>–</strong>sleeved shirts/pants for protection from sun/bugs<br />

Sweatshirt, sweatpants, socks in case you get chilled from being in water all day<br />

2-3 bathing suits (things don’t always dry over-night in the tropics)<br />

Well worn-in and comfortable walking shoes (sneakers or Teva-like sandals)<br />

Boat shoes (water shoes or bare feet are recommended on the boat)<br />

Rain gear or rain poncho<br />

Hat with wide brim to protect head from sun (very important!)<br />

Clothing/Footwear for Leisure<br />

One set of clothing to keep clean for day-off and end of expedition<br />

Field Supplies<br />

*****Polarized***** sunglasses with retaining strap<br />

Bound Field Journal & Copies of Journal Articles<br />

Small daypack/rucksack/backpack for inland trip<br />

Dry bag or heavy duty plastic “zip lock” bags for boat (for protecting equipment such as camera from sand,<br />

humidity, and water)<br />

Water bottle - 1 liter refillable, such as Nalgene<br />

Mask, snorkel, and fins if you want to participate in tasks requiring snorkeling, or for snorkeling during free<br />

time (we encourage you to invest in good quality mask and fins, not the department store variety)<br />

Dive skin or 1-2mm wetsuit<br />

Mosquito repellent (with Deet)<br />

Belize Field Course Briefing Page 13 4/4/<strong>2012</strong>


Oil or oil-based repellant (e.g. olive oil, AVON Skin-so-Soft Original Bath Oil, citronella oil repellent, Bit<br />

Blocker) for sandflies (oil creates a physical barrier from the sand flies; the researchers have found NOTHING<br />

except oil prevents the sand flies from feasting on you when/if the wind dies)<br />

A box of mosquito coils (commonly called “fish” in Belize) and a lighter for burning in your room at night to<br />

keep both mosquitoes and sandflies away<br />

Waterproof sunscreen with SPF 30 or higher<br />

Lip balm with SPF 30 or higher, also Blistex ointment if you are prone to fever blisters<br />

A notebook for your personal field notes and a couple of pencils (unlike ink pens, pencils continue to write even<br />

if they get wet)<br />

Personal Supplies<br />

Toiletries, such as a bar biodegradable soap, deodorant, toothpaste, toothbrush, shampoo, conditioner, and<br />

moisturizing lotion (no-fragrance types may help reduce the attraction of mosquitoes)<br />

Antibacterial wipes or lotion (good for “washing” hands while in the field)<br />

Personal medications, such as vitamins, prescription drugs, emergency allergy injections (Epi-kit), inhalants,<br />

motion sickness medication and over-the-counter antihistamines and anti-itch products<br />

Extra contact lenses and saline solution and/or spare glasses (you will need your reading glasses if you wear<br />

them)<br />

Miscellaneous<br />

Cash for snacks, gifts, tips, etc.<br />

Camera, film/memory cards, extra camera battery/battery charger, and computer cable for downloading images<br />

from digital cameras<br />

Laptop computer, power cord, power strip, flash drive or pen drive<br />

Cell phone with SMS Text Capability if you want to send and receive messages from home<br />

Optional Items<br />

Flashlight/torch or headlamp with extra batteries and extra bulb<br />

Earplugs (your roommates might snore!)<br />

Duct tape and roll of heavy string<br />

Very fine gauge mosquito net for hanging over your bunk (also to keep the sandflies out)<br />

Snack food<br />

Paperback books/field guides<br />

Small battery-operated fan with additional batteries<br />

Battery-operated tape/CD/MP3 player/recorder with headphones (NOT allowed on research boat)<br />

Binoculars<br />

Playing cards, board games, musical instruments, etc.<br />

Belize Field Course Briefing Page 14 4/4/<strong>2012</strong>


APPENDIX<br />

<strong>2012</strong> Course Flyer, Syllabus, Registration, & Policy Forms<br />

Field Journal Requirements<br />

Final Project Sample Fact Sheets<br />

Red Mangrove Root Crab<br />

Cushion Sea Star<br />

Fiddler Crab<br />

Patch Reef Habitats<br />

Brown Pelican<br />

Variegated Sea Urchin<br />

Red Fin Needlefish<br />

Fiddler and Hermit Crabs<br />

Cushion Sea Star<br />

Cushion Sea Star<br />

Magnificent Feather-Duster<br />

Marine Hermit Crab<br />

Variegated Sea Urchin<br />

Data Sheets for Manatee and Dolphin Research Project<br />

Faculty Curriculum Vitae<br />

Journal Club Readings<br />

Mortimer et al. 2007. Whose turtles are they, anyway? Molecular Ecology 16:<strong>17</strong>-18<br />

Blumenthal et al. 2009. Turtle groups or turtle soup: dispersal patterns of hawksbill turtles in<br />

the Caribbean. Molecular Ecology 18:4841-4853.<br />

Self-Sullivan. 2000. The Elusive Manatee: An ethological approach to understand behavior of<br />

the West Indian manatee in Belize.<br />

Self-Sullivan et al. 2003. Seasonal occurrence of male Antilllean manatees on the Belize<br />

Barrier Reef. Aquatic Mammals 29.3:342-354.<br />

Grimm. 2010. Is a Dolphin a Person? News Briefs, Science Magazine, 18-22 February 2010.<br />

Noren. 2008. Infant carrying behaviour in dolphins. Functional Ecology 22:284-288<br />

Miksis-Olds et al. 2007. Simulated vessel approaches elicit differential responses from<br />

manatees. Marine Mammal Science 23.3:629-649.<br />

Mann. 1999. Behavioral sampling methods for cetaceans: a review and critique. Marine<br />

Mammal Science 15.1:102-122.<br />

LaCommare et al. 2008. Distribution and habitat use of Antillean manatees in the Drowned<br />

Cayes, Belize. Aquatic Mammals 34.1:35-43.<br />

Kerr et al. 2005. Bottlenose dolphins in the Drowned Cayes, Belize. Caribbean Journal of<br />

Science 41.1:<strong>17</strong>2-<strong>17</strong>7.<br />

Jackson. 1997. Reefs since Columbus. Coral Reefs 16:S23-S32<br />

Harper & Schulte. 2005. Social interactions in captive female Florida manatees. Zoo Biology<br />

24:135-144.<br />

Belize Field Course Briefing Page 15 4/4/<strong>2012</strong>


Ecology, Behavior, and Conservation of Manatees & Dolphins<br />

A Unique Field Course in the Belize Barrier Reef Lagoon System<br />

<strong>2012</strong> Session III ~ 4 - <strong>17</strong> August <strong>2012</strong><br />

Lead Instructor & Principal Investigator: Caryn Self-Sullivan, Ph.D. 1,2<br />

Co-PI: Katie LaCommare, Ph.D. 2,3 | Visiting Faculty: TBA<br />

1 Nova Southeastern University, 2 <strong>Sirenian</strong> <strong>International</strong>, 3 Lansing Community College<br />

Want to be a Conservation Biologist, Behavioral Ecologist or Marine Mammalogist?<br />

Here's your chance to join our research team for two intense weeks of total immersion<br />

into the world of animal behavior, ecology & conservation, Antillean manatees,<br />

bottlenose dolphins, coral reefs, mangroves and seagrass beds in Belize!<br />

Course Overview: This is an experiential learning field course where you will live, work, and study<br />

from a marine science field station on a pristine, private island off the coast of Belize. Additionally, you<br />

will visit one or more Community Conservation Sites in Belize. Data collected during the course will<br />

contribute to our long-term manatee/dolphin research project. You will learn through a variety of<br />

learning activities, literature review and discussion, independent research projects, and actual field<br />

research. Be prepared to rise with the sun and spend 8-10 hours outdoors, including 3-4 hours on the<br />

water each day learning about the tropical Caribbean environment as we explore a maze of mangrove<br />

islands, seagrass beds, and coral patches searching for elusive manatees and charismatic dolphins.<br />

Location: Spanish Bay Conservation & Research Center at Hugh Parkey's Belize Adventure Lodge,<br />

http://belizeadventurelodge.com/ and Sarteneja Alliance for Conservation & Development,<br />

http://sartenejaconservation.org/. Passport required, immunizations as recommended by CDC<br />

Your Share of the Costs: US$2995 includes housing, meals, ground & water transfer fees, research &<br />

materials fees; DOES NOT include airfare, books, tips, and credit hours.<br />

Credit Hours: The course provides 100 experiential learning and lecture hours in the field, plus<br />

approximately 35 hours of pre-fielding research and preparation, is comparable to a 3 credit hour<br />

university course and meets the US DOE criteria in 34 CFR, §600.2. Deadlines: Early registration &<br />

and deposit due February 1st, <strong>2012</strong>; regular registration & deposit due March 1st, <strong>2012</strong>; balance due at<br />

least 60 days prior to field dates. Late payments and late registrations (if space available) incur a $100<br />

late fee.<br />

Information Contact: Caryn Self-Sullivan, Ph.D. | +1.540.287.8207 | cselfsullivan@sirenian.org<br />

=========================================================================<br />

cselfsullivan@sirenian.org<br />

540.287.8207<br />

cselfsullivan@sirenian.org<br />

540.287.8207<br />

cselfsullivan@sirenian.org<br />

540.287.8207<br />

cselfsullivan@sirenian.org<br />

540.287.8207<br />

cselfsullivan@sirenian.org<br />

540.287.8207<br />

cselfsullivan@sirenian.org<br />

540.287.8207<br />

cselfsullivan@sirenian.org<br />

540.287.8207<br />

cselfsullivan@sirenian.org<br />

540.287.8207<br />

cselfsullivan@sirenian.org<br />

540.287.8207<br />

cselfsullivan@sirenian.org<br />

540.287.8207<br />

cselfsullivan@sirenian.org<br />

540.287.8207<br />

cselfsullivan@sirenian.org<br />

540.287.8207


Ecology, Behavior, and Conservation of Manatees & Dolphins in Belize<br />

A Unique Field Course in the Drowned Cayes, Belize<br />

Summer <strong>2012</strong> SYLLABUS & TENTATIVE SCHEDULE<br />

Session III: ~ 4 <strong>–</strong> <strong>17</strong> August, <strong>2012</strong><br />

Visit us on Facebook: http://www.facebook.com/event.php?eid=370432825564<br />

Lead Instructor & PI<br />

Caryn Self-Sullivan PhD<br />

NSU & <strong>Sirenian</strong> <strong>International</strong><br />

Phone: 540.287.8207<br />

Email: cselfsullivan@sirenian.org<br />

or cs<strong>17</strong>33@nova.edu<br />

Co-PI: Katherine LaCommare PhD<br />

Lansing Community College &<br />

<strong>Sirenian</strong> <strong>International</strong><br />

Phone: 248-756-3985<br />

Email: kslacommare@gmail.com<br />

or katie.lacommare@umb.edu<br />

Want to be a Conservation Biologist, Behavioral Ecologist or Marine Mammalogist? Here's your chance to<br />

join our research team for two intense weeks of total immersion into the world of animal behavior, ecology &<br />

conservation, Antillean manatees, bottlenose dolphins, coral reefs, mangroves and seagrass beds in Belize!<br />

Course Overview: This is a total immersion, experiential learning, field course where you will live, work, and<br />

study from a marine science field station on a pristine, private island off the coast of Belize. Additionally, you<br />

will visit one or more Community Conservation Sites in Belize. Data collected during the course will contribute<br />

to our long-term manatee/dolphin research project established in 1998. You will learn through a variety of<br />

learning activities, literature review and discussion, independent research projects, and actual field research. Be<br />

prepared to rise with the sun and spend 8-10 hours outdoors, including 3-4 hours on the water each day learning<br />

about the tropical Caribbean environment as we explore a maze of mangrove islands, seagrass beds, and coral<br />

patches searching for elusive manatees and charismatic dolphins. Optional SCUBA dives are available as time<br />

and weather permits; NOTE: additional costs are involved for SCUBA diving.<br />

PRIMARY LOCATION: Spanish Bay Conservation & Research Center at Hugh Parkey's Belize Adventure<br />

Lodge, http://belizeadventurelodge.com/ (Passport required, immunizations as recommended by CDC)<br />

SECONDARY LOCATION(s): Corozal Bay Wildlife Sanctuary, co-managed by the Sarteneja Alliance for<br />

Conservation & Development, Sarteneja. Alternative 2ndary locations include: Swallow Caye Wildlife<br />

Sanctuary, co-managed by Friends of Swallow Caye, Caye Caulker; Gales Point Wildlife Sanctuary, Gales Point<br />

Manatee & Southern Lagoon; South Water Caye Marine Reserve (& Placencia Lagoon), co-managed by SEA<br />

Belize (formerly Friends of Nature & TASTE); Port Honduras Marine Reserve, co-managed by the Toledo<br />

Institute for Development and the Environment.<br />

COSTS: US$2995 includes housing, meals, ground & water transfer fees, research & materials fees; DOES NOT<br />

include airfare, books, tips, and credit hours.<br />

OPTIONAL CREDIT HOURS: The course provides 100 experiential learning and lecture hours in the field,<br />

plus approximately 35 hours of pre-fielding research and preparation; at least 45 of the 135 total hours include<br />

direct instruction by faculty. This is comparable to a 3 credit hour university course and meets the US DOE<br />

criteria in 34 CFR, §600.2. You must make arrangements IN ADVANCE with BOTH your advising faculty and<br />

Dr. Self-Sullivan for credit to be earned through your home university. Credit hour fees must be paid directly to<br />

your school and you must fulfill any study abroad requirements of your school. This course is divided into 4<br />

major components: short lectures and learning activities (~1 hour per day), independent reading and assignments<br />

(~2 hour per day), supervised data collection in the field (~3 hours per day), independent project development &<br />

implementation (~2 hours per day), presentation of pre-field research (~1 hour per day), and debate/group<br />

discussion of reading materials (~1 hour per day).<br />

DEADLINES: Early registration & and deposit due February 1st, <strong>2012</strong>; regular registration & deposit due March<br />

1st, <strong>2012</strong>; balance due at least 60 days prior to field dates. Late payments and late registrations (if space<br />

available) incur a $100 late fee. If you are registering through your home university, earlier deadlines may<br />

exist—see your academic advisor!<br />

Marine Mammal Ecology, Behavior, Conservation Field Course, Summer <strong>2012</strong>, Page 1 of 5, Revised 4/4/<strong>2012</strong>


MINIMUM / MAXIMUM CLASS SIZE: 6-24 students<br />

WEBPAGE: http://www.sirenian.org/<strong>2012</strong>FieldCourse.html<br />

Course Objectives: The objectives of this course are to introduce you to the basic concepts animal behavior,<br />

ecology, and conservation through participation in an ongoing research project, development and implementation<br />

of your own independent research project; research and presentation on Belizean culture or natural history,<br />

reading and thinking critically about peer reviewed literature related to course content, and participation in a<br />

Service-Learning Project.<br />

<strong>COURSE</strong> CONTENT (Readings, Lectures, Learning Activities) SUBJECT TO CHANGE DEPENDING ON<br />

FACULTY<br />

• Introduction to Trichechus manatus & Tursiops truncatus<br />

• Participation in long-term research project, results to date, questions for the future<br />

• Introduction to coral reef, mangrove & seagrass ecology<br />

• Introduction to Animal Behavior & Conservation<br />

RESEARCH ACTIVITES<br />

• Manatees and Dolphins: point transect scans, focal follows, reconnaissance surveys, photo-id<br />

• Habitat Sampling: seagrass identification to species level, percent cover, density, comparison of<br />

epibionts, core sampling, feeding scars.<br />

• Individual or team research project: select from list of approved study subjects and topics below<br />

Daily schedule: Our schedule is closely tied to sunrise/sunset and tides, with a tentative schedule, below. You<br />

will be actively engaged from 06:00 until 21:00, daily. Morning and afternoon learning activities and research<br />

will be rotated as necessary.<br />

Time Daily Schedule<br />

6:00 Observing Wildlife, Individual Project Data Collection<br />

7:30 Breakfast<br />

8:30 Learning Activities (alternating AM & PM)<br />

12:30 Lunch<br />

13:30 Field Research (alternating AM & PM)<br />

18:00 Journal Club & Happy Hour<br />

19:00 Dinner<br />

20:00 Belize Sea to Stars Presentations (20-40 minutes)<br />

22:00 Quiet Time in the Dorms (Reading/Writing OK)<br />

23:00 Curfew and Lights Out<br />

Books & Tools<br />

We recommend the following books, which are available from Amazon.com:<br />

The Florida Manatee: Biology and Conservation, by Roger L. Reep and Robert K. Bonde, University Press of<br />

Florida<br />

The Bottlenose Dolphin: Biology and Conservation, by John E. Reynolds, Samantha D. Eide, and Randall S.<br />

Wells, University Press of Florida<br />

A selection of primary literature is included in the Field Briefing Document. Students are required to print<br />

out and read this document prior to fielding AND bring it with you to Belize. We encourage you to have this<br />

done at a print shop, such as FedExOffice-Kinkos and bound with a 1” wire ring (~$35). There is also a library of<br />

books and archived journals, including Marine Mammalogy, Animal Behavior, and Conservation Biology<br />

journals, in our library onsite in Belize.<br />

Marine Mammal Ecology, Behavior, Conservation Field Course, Summer 2011 Page 2 of 5, Revised 4/4/<strong>2012</strong>


Communication: If you have any questions or concerns, it is your responsibility to communicate with us via<br />

email or phone prior to travel and in person while at the field site. As your professors, we are here to help you<br />

learn. But learning is your responsibility.<br />

CONTACT: Dr. Caryn Self-Sullivan, cselfsullivan@sirenian.org, US +1.540.287.8207, BZ +501.636.3849<br />

Student Expectations: You should expect us to be enthusiastic about the course material and to be available for<br />

one-on-one help and feedback. You should expect us to challenge you to think about how & why the course<br />

material is relevant to you and your career goals. You should expect us to respect you as a person, regardless of<br />

your academic performance.<br />

Our Expectations: We expect you to complete all pre-field requirements and readings by the due dates posted in<br />

the briefing. We expect you to give respect to everyone in Belize, especially at our field site. We expect you to<br />

fully participate in every course activity, including being present and on time for all meals. We expect you to<br />

demonstrate enthusiasm and respect for all living organisms in the tropical marine environment.<br />

Grading Structure: Your final grade is based on a 500 point scale. To determine your “letter” grade, divide<br />

points earned by possible points: A > 90%; B > 80%; C > 70%; D > 60%.<br />

• *Independent Research Project (100 pts): Select a topic from approved list** before fielding; do<br />

background research before departure; design and implement a short-term research project during field<br />

course; write up results in format of Fact Sheet (see examples from previous students).<br />

• Field Journal & Reflections (100 pts): Purchase and bring a bound notebook to be used as a field<br />

journal during the course. Make daily entries following guidelines presented in course; include a 101<br />

species list.<br />

• Participation in field and course activities (100 pts): Includes evaluation of your participation in all<br />

course activities, including discussions, quiz bowls, and personal essay “Why Marine Mammals?” Be on<br />

time and demonstrate enthusiasm for learning for full points.<br />

• *Paper discussion and leading (100 pts): Read all papers prior to coming into the field; review in field<br />

and participate in all discussions; present and lead the discussion as assigned.<br />

• Service-Learning (50 pts): Either as a class or in mini-groups we will contribute the improvement of the<br />

island or regional environment. Projects could include picking up garbage in the mangroves, creating<br />

signage material for education, helping in the Swallow Caye refuge, participating in stranding or<br />

necropsy, giving a presentation at a local school or civic group, giving a talk to island staff.<br />

• *Belize from Sea to Stars (50 pts): Each student will prepare a presentation before arrival on some<br />

aspect of Belize. Topics could include the following**: Art, Astronomy, Culture, Economics, Ecology,<br />

Geology, History, Languages, Music, Politics and others. After dinner, a student will give an oral<br />

presentation on their topic. They may prepare handouts but this will not be a power point talk. For full<br />

points, there should be some interactive activity such as a quiz bowl, role playing, or crossword puzzle.<br />

These should be informative and interesting to a general audience. The island staff will be invited to join<br />

us when this is feasible.<br />

* Requires pre-course reading, research, and communication **See suggested topics at end of syllabus<br />

Missed and Late Assignments: Each Assignment (pre-fielding and in the field) will have a firm due date and<br />

time. If you miss the due date/time, you may turn in your assignment late. However, your grade will be reduced<br />

by 10% per day until the last day of the course. Any assignments not completed by the last day of the course will<br />

receive zero credit.<br />

Academic Dishonesty & Plagiarism: Students found violating the conditions of academic honesty will<br />

receive an F in the course. DO NOT copy & paste text from the Internet or other sources into your<br />

assignments; DO NOT copy work from your peers; DO NOT paraphrase someone's ideas without giving<br />

appropriate credit to the source. Plagiarism is a serious offense and will be treated as such. For more<br />

information on plagiarism see http://en.wikipedia.org/wiki/Plagiarism.<br />

Marine Mammal Ecology, Behavior, Conservation Field Course, Summer 2011 Page 3 of 5, Revised 4/4/<strong>2012</strong>


<strong>COURSE</strong> FEE & ADDITIONAL INFORMATION<br />

The course fee of $2995 includes all transportation, meals, and accommodations from your arrival at the PWG<br />

airport in Belize on the first day of the course to your departure on the last day of the course. Additionally, you<br />

are responsible for tips, insurance, credit hours, and round-trip airfare to Belize (airport code BZE).<br />

EARLY REGISTRATION DEADLINE: 1 February <strong>2012</strong><br />

REGULAR REGISTRATION DEADLINE: 1 March <strong>2012</strong><br />

LATE PAYMENTS OR LATE REGISTRATION: $100 Late Fee<br />

To register for course, complete, sign, and send these forms: http://sirenian.org/<strong>2012</strong>BelizeRegistration.pdf<br />

http://sirenian.org/<strong>2012</strong>BelizeSyllabus.pdf<br />

http://sirenian.org/<strong>2012</strong>BelizePolicy.pdf<br />

Download, print, bind, READ, and bring to Belize: http://sirenian.org/<strong>2012</strong>BelizeBriefing.pdf<br />

Recruit a friend: http://sirenian.org/<strong>2012</strong>BelizeFlyer.pdf<br />

For more information on the course, please email: cselfsullivan@sirenian.org<br />

For more information on the facilities, please visit: http://belizeadventurelodge.com<br />

http://sartenejaconservation.org/<br />

PRINT STUDENT NAME: ___________________________________________________________________<br />

I HEREBY CERTIFY THAT I HAVE READ AND FULLY UNDERSTAND THIS SYLLABUS AND THE<br />

ADDITIONAL REQUIRED DOCUMENTS LISTED ABOVE.<br />

STUDENT SIGNATURE: ____________________________________________________________________<br />

MY CHOICES FOR BELIZE SEA TO STARS AND INDEPENDENT RESEARCH TOPICS ARE:<br />

RESEARCH 1 ST CHOICE:<br />

RESEARCH 2 ND CHOICE:<br />

RESEARCH 3 RD CHOICE:<br />

BELIZE 1 ST CHOICE:<br />

BELIZE 2 ND CHOICE:<br />

BELIZE 3 RD CHOICE:<br />

Questions, Concerns, Feedback, Comments:<br />

Marine Mammal Ecology, Behavior, Conservation Field Course, Summer 2011 Page 4 of 5, Revised 4/4/<strong>2012</strong>


Topics for Independent Research Project (100 pts): Select a species and/or ecosystem from list below & get<br />

approved by Dr. C BEFORE fielding; do background research (primary literature search) in advance so<br />

that you can design and implement a short-term research project during field course; write up results in<br />

format of Fact Sheet (see examples from previous students).<br />

Seagrass ecosystem: Abundant seagrass beds dominated by Thalassia testudinum, with Syringodium filiform,<br />

Halodule spp., surround the caye, providing an excellent environment for snorkel based investigations. Potential<br />

projects include ecology or behavior of various organisms within these systems. Suggested species include sea<br />

stars (Oreaster reticulates), sea urchins (Lytechinus spp.), grunts (Haemulon spp.), needle fish (Strongylura spp.).<br />

Shallow subtidal mangrove peat banks surround the caye and have well-developed communities of fleshy algae<br />

(e.g., Caulerpa spp., Dictyota spp.) and calcified algae (e.g., Halimeda spp., Jania spp.) with sparse T.<br />

testudinum, rose corals (Manicina areolata), and corkscrew anemones (Bartholomea annulata) with symbiotic<br />

cleaner shrimp (Periclimenes spp.).<br />

Mangrove ecosystem: The caye is dominated by large perimeter and dwarf central red mangroves (Rhizophora<br />

mangle) and 2 significant black mangrove (Avicennia germinans) forests, with a scattering of white mangroves<br />

(Laguncularia racemosa), a few buttonwoods (Conocarpus erectus), and planted coconut palms, providing<br />

abundant habitat for hermit crabs (Coenobita clypeatus) and fiddler crabs (Uca spp.); we also sight mangrove<br />

crabs (Aratus pisonii), snails (Littoraria angulifera), mangrove root crabs (Goniopsis cruentata), and ghost crabs<br />

(Ucides cordatus), mangrove rivulus fish (Rivulus marmoratus), and giant termite nests (Nasutitermes spp.),<br />

unidentified bat species locally called rat bats that are thought to roost in the coconut palms during the day but can<br />

be observed foraging on insects at dusk and dawn, and many birds including the mangrove warbler, golden<br />

fronted woodpecker, osprey, common black hawk, egrets, and herons (see bird list). The central pond of on this<br />

caye is strongly influenced by diurnal tides, fluctuating from a mud flat at low tide to about 30cm of water at high<br />

tide and would make a good independent variable for natural experiments.<br />

Belize from Sea to Stars (50 pts): Each student will prepare a presentation before arrival on some aspect of<br />

Belize. Topics could include the following: Art, Astronomy, Culture, Economics, Ecology, Geology, History,<br />

Languages, Music, Politics and others. After dinner each night, one or more students will give oral presentations<br />

on their topic. This is NOT a Power Point presentation! Ideally, there should be some interactive nature such<br />

as stargazing, quiz bowl, role playing, or crossword puzzle. These should be approximately 20-40 minutes long,<br />

informative and interesting to a general audience. The island staff will be invited to join us when this is feasible.<br />

Once you have selected your topic, you must register it with Dr. C to prevent replication of topics by students.<br />

For some interesting issues in Belize search the following topics in association with Belize:<br />

Lethal Yellowing Disease, Coral Bleaching<br />

Business Ventures: Chalillo Dam, BEL, Fortis; Cruise<br />

Ship Industry explosion; sugar and banana industries<br />

Glow Worms and Bioluminescence (Odontosyllis<br />

luminosa, Dr. Gary Gaston); Pica Pica and Thimble<br />

Jellies<br />

Parliamentary Government: UDP vs. PUP<br />

Meteor Showers, the Southern Cross and other<br />

Astronomical Views<br />

Culture: Maya, Garifuna, Kriole/Creole, Mestizo,<br />

Mennonites<br />

Geography: 5 Oceanic Ridges and the Atolls<br />

Local NGOS: BACONGO, TIDE, TASTE, Friends of<br />

Nature, Belize Audubon, Wildtracks, Sharon Matola and<br />

the Belize Zoo<br />

Some Famous People from Belize: Baymen of Belize and<br />

the <strong>17</strong>98 Battle of St. George’s Caye, Baron Bliss, Andy<br />

Palacio, George C. Price, Marion Jones, Jamal Shyne<br />

Barrow, Evan X Hyde, Zee Edgell, Milt Palacio, Said<br />

Musa, Dean Barrow<br />

Conservation in Belize: Protected Areas, Hicatee Turtles,<br />

Coral Reefs, Harpy Eagle, etc.<br />

Xaté (sha-tay): leaves from three Chamaedorea palm<br />

species (C. elegans, C. oblongata and C. ernesti-augustii)<br />

used in the floral industry<br />

Belize-Guatemala Border Dispute<br />

Music: Stonetree Records<br />

Destinations: Caye Caulker, San Pedro, Placencia,<br />

Orange Walk, Punta Gorda, Dangriga, Belmopan,<br />

Corozal, Bermudian Landing, San Ignacio, Benque Viejo,<br />

Hopkins, Crooked Tree, Sarteneja<br />

Maya Archeological Sites: Xunantunich, Cahal Pech,<br />

Lamani, Altun Ha, Caracol, El Pilar, Lubaantun, Nim Li<br />

Punit, Actun Tunichil Muknal, Cerros, Tikal (Guatemala)<br />

Food: Sere, Hudut, Escabeche, Rice and Beans, Beans<br />

and Rice, Cowfoot Soup, Johnny Cakes, Creole Bread,<br />

Powder Buns, Black Cake, Dukunu, Meat Pies, Relleno,<br />

Stew Chicken, Garnaches, Salbutes, Panades, Tambrand<br />

Candies, Gibnut, Bamboo Chicken, Marie Sharp’s, Boilup,<br />

etc.<br />

Invasives: Lionfish Pterois volitans, ?P. miles<br />

Marine Mammal Ecology, Behavior, Conservation Field Course, Summer 2011 Page 5 of 5, Revised 4/4/<strong>2012</strong>


Ecology, Behavior & Conservation of Manatees and Dolphins<br />

A Unique Field Course in the Drowned Cayes, Belize<br />

Session III: August 4-<strong>17</strong> <strong>2012</strong><br />

Step 1: PRINT & COMPLETE THIS FORM - ALL PRICES ARE IN $US<br />

Step 2: Mail, Fax, or Scan & Email the SIGNED registration, syllabus, transcripts & policy forms<br />

Step 3: I will send you and invoice, which you can pay online or by mail<br />

Mailing Address, Email Address, and Fax:<br />

Caryn Self-Sullivan, 200 Stonewall Drive, Fredericksburg, VA 22401-2110<br />

Phone: 540.287.8207 | Fax: 888.371.4998 (Country Code 001)<br />

Email: Caryn Self-Sullivan (cselfsullivan@sirenian.org)<br />

Name:<br />

Mailing Address:<br />

City, State, Postal Code, Country:<br />

Email Address(es):<br />

Cell & Home Phone #s:<br />

(cell) (home)<br />

Why do you want to participate in this course? You do NOT need to be currently enrolled in school, however describe your academic status & highest<br />

degree. Don't forget to send unofficial copies of your transcripts and answer questions on page 2:<br />

Openings available only in Session III Fee<br />

Session I: 26 May - 8 June <strong>2012</strong> $2,995<br />

Session II: 16 June - 29 June <strong>2012</strong> $2,995<br />

Session III: August 4-<strong>17</strong> <strong>2012</strong> $2,995<br />

Note: If you register for more than one session, you will be responsible for expenses b/t sessions!<br />

Total Cost (excludes airfare, credit hours & tips)<br />

DEPOSIT (Minimum $500) Due now; I will invoice you upon receipt of registration!<br />

Low-income Country Scholarship* $500<br />

Early Registration & Recruiting Discounts* (Max $100)<br />

Late Payment / Late Registration Fee $100<br />

Balance due at least 60 days prior to fielding<br />

* Attach a 1500 word essay describing how this opportunity will enhance both your personal career goals AND<br />

sirenian research, educational outreach, and conservation in your home country<br />

** $100 discount if you register and submit $500 deposit by February 1st; $25 discount each for you and a friend, for<br />

each friend you recruit, provided you and your friend(s) register and submit $500 deposit by March 1st; $25 discount if<br />

you register and pay in full by March 1st ; Mix and Match discounts for up to a maximum total discount of $100.<br />

By signing below, I hereby promise to pay the price indicated above subject to qualified discounts<br />

and the refund policy as stated in the Syllabus. I understand that I will be invoiced for the deposit<br />

immediately, and balance is due at least 60 days prior to fielding. Invoices will be sent via PayPal<br />

from "MIB: Research & Educational Expeditions." A $100 fee will be assessed for returned checks<br />

or chargebacks.<br />

Signature ____________________________________________________________________<br />

Comments or Questions:


Please provide the following information about yourself. Use additional pages if necessary!<br />

* Gender:<br />

* Date of Birth:<br />

* Citizenship:<br />

male female (circle or highligh)<br />

** Passport Country and Number:<br />

** DAN Preferred Membership Number:<br />

*REQUIRED at time of application **REQUIRED at least 60 days prior to fielding<br />

Are you currently enrolled in school? Please give details about your status and program of study. Attach additional<br />

pages if necessary!<br />

Do you want/need credit for this course? How many hours?<br />

If yes, please provide contact information for your academic advisor:<br />

Name:<br />

Department:<br />

University:<br />

Phone:<br />

Email:<br />

Are you SCUBA Certified? YES NO<br />

If yes, provide a copy of your Certification Card and DAN Card.<br />

Tell us about other field courses or field work you have done:<br />

What should we know about your history (allergies, dietary restrictions, psychological conditions, physical or chronic<br />

conditions, etc.) that might affect your ability to participate in this course?<br />

Why did you chose this course? What are you expectations? Do you have any specific background, skills, or talents?<br />

Attach additional pages if necessary!<br />

Tell us about your experiences working with teams, with different cultures, or other situations that demonstrate your<br />

adaptability. Use additional pages if necessary!<br />

Circle or highlight your current swimming ability:<br />

non-swimmer recreational swimmer strong swimmer life-saving certificate<br />

Circle or highlight your comfort level in the ocean:<br />

not at all comfortable in calm seas comfortable in choppy seas<br />

Describe any life saving training you have done, include dates (Red Cross, First Aid, CPR, EMT, etc.)?<br />

What kind of snorkel, SCUBA, and/or boating experience (power, sail, kayak, etc.) do you have?


<strong>2012</strong> Field Course Policy & Liability Release Form<br />

Ecology, Behavior & Conservation of Manatees & Dolphins in Belize<br />

August 4-<strong>17</strong>, <strong>2012</strong><br />

Field Class Policy & Liability Release<br />

COMPLETE, SIGN, INITIAL, AND RETURN BY REGISTRATION DEADLINE<br />

As your instructors, it is our responsibility to ensure that students, staff, colleagues, and associates are familiar<br />

with and understand field course policies. As a student participant, you are required to read, fill out, and sign<br />

this document where indicated. Please feel free to contact me if you have any questions about our Field Class<br />

Policy & Liability Release Form.<br />

Return this signed document via mail, FAX, or Email at the same time you submit your Registration and Signed<br />

Syllabus (page 4 only). Also, be sure to send copies of your Passport, Dive Card, DAN Card, and Flight<br />

Itinerary ASAP, but not later than 60 days prior to fielding, to: Caryn Self-Sullivan, 200 Stonewall Drive,<br />

Fredericksburg, VA 22401-2110; Email: cselfsullivan@sirenian.org; Phone: 540.287.8207; FAX: 888.371.4998<br />

As Principal Investigator and Lead Instructor, I am responsible for:<br />

• ensuring policies are followed;<br />

• ensuring that risks are minimized to the best of my ability;<br />

• ensuring that students or staff are removed from the course if they, in my sole judgment, are in violation of these<br />

policies and/or present a danger to the safety and/or success of the course.<br />

Should a student become seriously injured or leave the course early for any reason, I will inform the student's emergency<br />

contact that they have left the course and why. Your signature below authorizes me to do so! Please complete your<br />

emergency contact information:<br />

In case of emergency or my removal from this course, please contact the following person(s):<br />

Name(s): __________________________________________________________________________________<br />

Address: __________________________________________________________________________________<br />

Phone(s): __________________________________________________________________________________<br />

Email(s): __________________________________________________________________________________<br />

Relationship to Student: ______________________________________________________________________<br />

Student Signature: ___________________________________________________________________________<br />

CONSENT TO INHERENT RISKS<br />

Some of the characteristics that make a field course attractive to students may also put the student or their property at risk.<br />

To reduce this risk, all services from arrival to and departure from Philip S. W. Goldson <strong>International</strong> Airport (PGIA,<br />

airport code BZE) are provided by Hugh Parkey's Belize Dive Connection (BDC), which has over 20 years of experience in<br />

Belize and also has liability insurance in place.<br />

Each student agrees to purchase international medical travel insurance in the form of a DAN Preferred Membership<br />

(Divers Alert Network) to cover his/her expenses in case of an emergency. This policy can be purchased online at<br />

http://www.diversalertnetwork.org/ under the MEMBERSHIP tab. Sign up as an Individual (not a student) and select the<br />

BEST: Preferred Membership Plan ($110). Note: you do not have to be a certified diver to purchase DAN Preferred<br />

Membership. Some of the potential risks inherent to this course are listed below. Initial Here: Yes ____ No ____<br />

Climate: Students must be prepared to spend long days in hot, humid, and wet conditions. The tropical sun is very strong.<br />

Risks include dehydration, sunburn, and other heat related illnesses. Students agree to take precautions as advised by<br />

instructor and staff to prevent illness. Students agree to notify instructor immediately if they are feeling ill.<br />

Insects: Sand flies and mosquitoes can be problematic on the caye. Sand flies are believed to be a vector for leishmaniasis<br />

in some regions of Belize, but this has not been reported for the cayes. Mosquitoes may transmit a number of diseases (see<br />

Diseases below). Botflies (Dermatobia hominis) are also reported in Belize, and mosquitoes may transmit their larvae to<br />

human hosts where the larvae will grow and develop. This is not life threatening, but can be painful and unpleasant.<br />

Students agree to keep instructor informed of insect bites.<br />

Page 1 of 4 Updated by CSS on 4/4/<strong>2012</strong>


<strong>2012</strong> Field Course Policy & Liability Release Form<br />

Marine life: Fire corals, several species of stinging jellyfish, sharks, sea urchins, lionfish, and other potentially dangerous<br />

marine organisms can be found in the study area. These can inflict painful and occasionally severe stings or bites, which<br />

may become infected if untreated. Those with a serious allergy to bee or wasp stings may have a similarly dangerous<br />

reaction to corals and jellyfish. The best prevention is to avoid touching unfamiliar plants & animals. Students agree to<br />

notify instructor immediately if they are injured as a result of an interaction with marine life. Students with known allergies<br />

must notify instructor on this form and bring their own medications and an 'epipen' as prescribed by their physician.<br />

Snorkeling: All the inherent risks of snorkeling are obviously present, including the effects of environmental conditions,<br />

marine life and other risks specific to your own physical/medical history. Snorkeling is optional. Students who chose to<br />

snorkel should bring their own gear (mask, snorkel, fins, wet suit or skin) and know how to snorkel safely without<br />

hyperventilating. A snorkel "check-out" by the instructor will be conducted on day 2 of the course.<br />

Boats: We will be aboard a small fiberglass boat for all fieldwork. All boats should have ladders and a bimini cover for<br />

sun protection; however, in some instances we may use a boat that lacks a ladder or shade. Deck surfaces of boats will<br />

become slippery and may place you at risk of slips, falls, and injuries that result from these accidents. All boats are operated<br />

by a licensed captain, employed by BDC. Students are not allowed to operate boats during the expedition.<br />

Disease: Diseases found in tropical regions include malaria, dengue fever, filariasis, leishmaniasis, onchocerciasis,<br />

trypanosomiasis (Chaga’s disease), schistosomiasis, leptospirosis, rabies, brucellosis, hepatitis, and typhoid. Most diseases<br />

are preventable with basic safety cautions. Students certify that they have sought professional advice from a physician,<br />

local health department, or international travel clinic regarding immunizations against diseases.<br />

Driving: Driving conditions in Belize are considered poor by US standards and pose inherent risks. All land transfers will<br />

be done by licensed drivers, employed by Hugh Parkey's Belize Adventures. Students will not be permitted to drive during<br />

the expedition.<br />

INTELLECTUAL PROPERTY RIGHTS<br />

Students may share photos, videos, and stories of the expedition with family, friends, local media, and in public forum for<br />

educational outreach purposes.<br />

However, all information (data and images) shared or gathered during the expedition remains the intellectual property of PI<br />

Caryn Self-Sullivan (CSS) and Co-PI Katherine S. LaCommare (KSL). Students are strictly forbidden from co-opting or<br />

plagiarism of data, images or information gathered during an expedition for use in a scientific thesis, masters or PhD work,<br />

for profit, or for the academic or business use of a third party without the express written permission of CSS & KSL.<br />

Students are aware that data gathered during the interviewing of local people becomes the intellectual property of CSS &<br />

KSL.<br />

Students are sometimes required to submit a written report reflecting what they have learned on a project, sometimes as a<br />

step toward receiving credit at their home institutions. Permission is hereby granted to students to use data and images for<br />

this purpose.<br />

LIFESTYLE CHOICES<br />

Our course does not discriminate on the basis of race, religion, ethnicity, or sexual orientation; and we respect students'<br />

rights to privacy. However, students should be warned that their lifestyle decisions may offend or clash with the<br />

sensibilities of local residents in our area of operations, or potentially violate local laws. Further, certain lifestyle decisions<br />

and behavior that impacts fellow students or the instructor may result in an uncomfortable, hostile and/or unproductive<br />

work environment.<br />

To ensure enjoyable and productive work conditions and smooth relations with local peoples, we have established the<br />

following code of conduct. Beyond practicing cultural sensitivity and showing common courtesy, please be mindful of the<br />

following limitations, and please remind and encourage all participants to practice sensitivity, courtesy, and compliance<br />

with these policies.<br />

Fraternization: Instructors, staff, colleagues, and associates are prohibited from becoming intimately involved with<br />

students during the duration of the field course. Romantic relationships between students that may otherwise seem<br />

permissible often create an unpleasant and/or unproductive work environment and are therefore strongly discouraged for<br />

the duration of the field course.<br />

Sexual Harassment: The inter-personal dynamics that exist between field course instructors, staff, and students are<br />

considered student-teacher relationships. Therefore, please be aware of the following policies.<br />

• Sexual harassment of students by the instructor or staff is prohibited.<br />

• Likewise, sexual harassment of the instructor, staff, other students, or local people by students is also prohibited.<br />

Page 2 of 4 Updated by CSS on 4/4/<strong>2012</strong>


<strong>2012</strong> Field Course Policy & Liability Release Form<br />

Sexual harassment infringes on an individual’s right to an environment free from unsolicited and unwelcome sexual<br />

overtones of conduct either verbal or physical. Sexual harassment does not mean occasional compliments of a socially<br />

acceptable nature. Sexual harassment refers to conduct which is offensive, which harms morale, or which interferes with<br />

the effectiveness of field courses and research expeditions. Such conduct is prohibited. Lewd or vulgar remarks,<br />

suggestive comments, displaying derogatory posters, cartoons or drawings, pressure for dates or sexual favors and<br />

unacceptable physical contact or exposure are examples of what can constitute harassment. It is important to realize that<br />

what may not be offensive to you, may be offensive to other students, the local population, staff members, or an instructor.<br />

Any individual who feels subjected to sexual harassment or has any knowledge of such behavior should report it at once to<br />

his or her instructor or staff members. All reports of sexual harassment will be handled with discretion and will be<br />

promptly and thoroughly investigated. Any student who is found to have engaged in conduct constituting sexual<br />

harassment will be removed from the course immediately. Any expenses involved in an early departure from the course are<br />

the responsibility of the student.<br />

Drugs: The manufacture, possession, use, purchase and/or sale of illegal drugs as defined by the United States Code and/or<br />

Belize Law is strictly forbidden while participating in this field course. Prescription drugs may only be purchased and used<br />

by the individual indicated on the prescription, in keeping with the intended use guidelines.<br />

Alcohol: Instructors and staff members are in a position of leadership and must therefore always be prepared to make<br />

responsible, timely decisions in any situation that may arise. If you decide to consume alcohol after course work and<br />

research has concluded for the day, please exercise your best judgment. Please respect local laws and customs.<br />

Intoxication can jeopardize your own safety, in addition to that of other students and staff.<br />

Students who are US Citizens must be a minimum of twenty-one (21) years old to consume alcohol. Be prepared to show<br />

your identification to the HP staff prior to the purchase of alcoholic drinks. Per day limits on alcohol consumption are in<br />

place during the course. In addition, restrictions on the use, possession, sale, or purchase of alcohol are solely at the<br />

discretion of the instructors. Local statutes, customs, practices, ordinances, and regulations with regard to the use,<br />

possession, sale, or purchase of alcohol are applicable to all students, instructors, and staff participating in this course.<br />

The instructor has the discretion to remove individuals from the course who consume alcohol in a time and manner that<br />

endangers the safety and/or productivity of other students, staff, the course, or any expedition.<br />

RECREATIONAL DAYS<br />

As an instructor and part-time resident of Belize since 1998, we (CSS & KSL) are familiar with local conditions and<br />

activities that may place the student at risk. Depending on the size of the student body, students may be given options for<br />

recreational day activities, but all activities must be conducted through BDC or their authorized agents. Potentially<br />

dangerous recreational day activities will not be allowed, as they may jeopardize your safety, the safety and well being of<br />

students, and ultimately, the success of this long-term research project.<br />

Students will occasionally have unsupervised free time before and after classes & discussions/field trips/research sessions.<br />

Activities undertaken during unsupervised time will be at your own risk.<br />

IN THE EVENT OF AN EMERGENCY: GOOD SAMARITAN ACTIONS<br />

In the event of emergencies, instructors, staff, and students must make certain judgments. While we have made an effort to<br />

ensure that qualified people make the most informed decisions possible, occasionally First Aid must be administered and<br />

other immediate steps taken by expedition participants who are not officially certified to make these decisions. We have<br />

safety protocols and emergency procedures in place. However, in rare, unforeseeable emergency situations, you are not<br />

restricted from exercising your best judgment with regard to your own safety. We do not restrict “good Samaritan”<br />

actions, or actions taken to assist fellow participants during emergency situations in the field. In the event of an unforeseen<br />

contingency that has not been provided for in BDC's emergency plan, we will take all reasonable steps to assist students<br />

and each other, with the understanding that we do not expect you to take steps that unreasonably jeopardize your own<br />

safety or that of others.<br />

STUDENTS AND DRIVING<br />

Students are not allowed to drive project vehicles or boats during the field course. Students may not transport themselves,<br />

other students, staff, or equipment in a vehicle of any kind, including their own. Students may only drive a project vehicle<br />

or boat, transport other students or perform errands in the field in case of an emergency and when no authorized drivers are<br />

available.<br />

Page 3 of 4 Updated by CSS on 4/4/<strong>2012</strong>


SCUBA, DIET, ALLERGIES<br />

<strong>2012</strong> Field Course Policy & Liability Release Form<br />

I (AM) or (AM NOT) (circle one) certified to dive using SCUBA equipment by _________________________<br />

_________________ (certifying organization). If certified, my certification level is ______________________;<br />

my certification number is ____________________; and I have attached a copy of my Dive Card to this form.<br />

Dietary Preferences (circle one): Vegan Vegetarian Carnivore Omnivore<br />

Allergies (circle one): NONE FOOD (list below) DRUGS (list below) OTHER (list below)<br />

REFUND POLICY<br />

I hereby acknowledge that if I cancel my registration before May 1 st , I will receive 100% of my deposit less a<br />

$50 service fee to cover bank and processing costs. If I cancel between May 1 st and June 1 st , I will forfeit my<br />

$500 deposit, but will receive a full refund of any balance paid above the deposit, less a $50 service fee. No<br />

refunds will be given after June 1 st . However, if the course is canceled by the instructor for any reason, I<br />

understand that I will receive a full refund of all payments made. I agree that I will NOT purchase my airfare<br />

until the instructor has confirmed that the course has made minimum enrollment. Initial Here: Yes ____ No ____<br />

PERMISSION TO RELEASE YOUR IMAGE AND INFORMATION<br />

I hereby consent to Caryn Self-Sullivan, the Hugh Parkey Foundation, and Belize Dive Connections releasing or<br />

using photographs of me taken during my field course & research expedition. Initial Here: Yes ____ No ____<br />

I hereby consent to Caryn Self-Sullivan releasing or using information gathered about me in the course of my<br />

field course & research expedition, except for contact and medical information and/or information that is<br />

otherwise agreed to be kept confidential. Initial Here: Yes ____ No ____<br />

ASSENT TO POLICIES<br />

I have read and understand the policies, rights, and responsibilities expressed in this document. I accept the<br />

policies above as a condition of participating on this field course and I accept the consequences described above<br />

for gross negligence or serious violations of course policies.<br />

RELEASE OF LIABILITY<br />

Student Signature: _____________________________________Date: ____/____/____<br />

I understand the potential hazards and risks attendant to the field course that is described in the syllabus, course<br />

briefing, and this policy document. By signing below I agree to participate in the field course at my own risk. I<br />

declare that I am in good health. I declare that I, and my heirs, in consideration of my participation in the<br />

Ecology, Behavior & Conservation of Manatees, Dolphins, Turtles field course from 18 th to 31 st May 2011 in<br />

Belize, hereby release Caryn Self-Sullivan, course instructors, and visiting scientists from any and all liability<br />

for damage to or loss of personal property, sickness or injury from whatever source, legal entanglements,<br />

imprisonment, death, or loss of money, which might occur while participating in this course. I am aware of the<br />

risks of participation, which include, but are not limited to, the possibility of injury or death. I hereby state that I<br />

am in sufficient physical condition to accept a normal level of physical activity during long-term exposure to the<br />

sun and heat. I understand that participation in this program is strictly voluntary and I freely chose to participate.<br />

I understand that the course does not provide medical coverage for me. I verify that I have purchased travel<br />

insurance equivalent to DAN Preferred Membership and that I will be responsible for any medical costs I incur<br />

as a result of my participation.<br />

However, I do not release Caryn Self-Sullivan, course instructors, and visiting scientists from liability on<br />

account of any injury, loss, or damage to me directly caused by the gross negligence or wanton or reckless<br />

misconduct of Caryn Self-Sullivan, course instructors, and visiting scientists.<br />

Student Signature: _____________________________________Date: ____/____/____<br />

Page 4 of 4 Updated by CSS on 4/4/<strong>2012</strong>


FIELD JOURNALS<br />

You are required to keep a permanently bound field journal or notebook (no loose leaf binders) in<br />

addition to having a notebook for your course notes. We encourage you to make daily entries following<br />

guidelines presented here and to include a “101 Species List” in the back of the journal. Your goal is to<br />

identify at least 101 organisms to the species level during the course including the 3 most prominent sea<br />

grass species, 4 mangrove species, common birds, fish, corals, reptiles, and insects. Your list should<br />

include common name, Genus and Specific Epithet, Phylum, identification keys, and a sketch. You will<br />

use information from your journal to complete your final exam, which will be in the form of a scavenger<br />

hunt!<br />

Memories are ephemeral and perceptions change with time, thus you should have your field journal with<br />

you at all times during to document your field experience, journal readings, and research efforts. Some<br />

researchers tend to write rough field notes in their journal while in the field, and then write a more<br />

complete description and expand on their observations each night after the fieldwork--in the same journal.<br />

This method has the advantage of allowing you to recognize what type of information you are omitting<br />

from your field notes and make sure you collect those data in the future.<br />

Your journals must be neat, legible and have appropriate spelling, etc. We recommend using a pencil or<br />

permanent ink pen (i.e., Pigma pens, Rite-in-the-Rain, etc.) since we may be in situations in which the<br />

pages of your journal will get wet. Most local stores that carry office supplies also have journals that are<br />

waterproof. While that type of journal is not required, it is encouraged.<br />

Drawings and sketches are an invaluable addition to your journal. Even the lousiest artist (like me :-) can<br />

make a decent field drawing in about 2 minutes. Start with the sea grasses and mangroves, then move on<br />

to the animals! A portion of your final grade will be based on the completeness, legibility, and accuracy<br />

of your journal. Also, do not erase incorrect entries but strike through them with a single line and write<br />

any correction above or below the original entry.<br />

A good field journal includes time, date, place, conditions, guests, professors, and other information at the<br />

heading of each entry. Entries should include information about each site and details about field<br />

techniques as well as information about your activities and data collection during your research project.<br />

In addition, entries should contain reflective thought about how you feel about the experience.<br />

Finally, you should always include contact information for anyone you meet in the class or the field as it<br />

may assist you in future endeavors. We are happy to review your journal at any time and it will be<br />

collected briefly and graded on the last day of class.<br />

Things that must be included in your journals:<br />

1. Your name, phone number, email and permanent address<br />

2. The names and contact information for your research partner<br />

3. Dates, times, weather conditions, and location of all field activities that are of sufficient detail to<br />

be replicated<br />

4. Name and contact information for every researcher or staff member that provides you information<br />

(including Belizeans!)<br />

5. Descriptions of field trips with detail on methodologies learned that would allow replication<br />

directly from your notebook<br />

6. Brief summaries of your daily activities<br />

7. All data and summaries of your thoughts regarding your individual research projects<br />

8. Your thoughts and reflections about each of the field experience<br />

9. Complete citations and notes from any sources (readings, etc.) so you can reference them in the<br />

future<br />

By J. Young (2010), Revised by C. Self-Sullivan (2011)


Red Mangrove Root Crab<br />

Goniopsis cruentata<br />

Hana Bucholz and Meagan Wise<br />

Edited by Caryn Self-Sullivan, Ph.D.<br />

Ecology & Behavior of Manatees & Dolphins in<br />

Belize, Class of 2009 (contact: caryns@sirenian.org)<br />

Taxonomy<br />

Kingdom Animalia<br />

Phylum Arthropoda<br />

Class Crustacea<br />

Order Decapoda<br />

Family Grapsidae<br />

Genus Goniopsis<br />

Species Goniopsis cruentata<br />

Ecology<br />

One of the more elusive crab species on Spanish<br />

Lookout Caye is the mangrove root crab<br />

(Goniopsis cruentata). Also known simply as<br />

the root crab, this brightly colored crab (purple,<br />

red and orange with white spots) lives among<br />

the red mangrove roots and branches. During<br />

low tide, you can find it beneath the red<br />

mangrove roots (Rhizophora mangle) and<br />

pimento stakes along the island’s central<br />

causeway. According to Feller (1996), these<br />

crabs have been observed feeding on mangrove<br />

propagules, insects, and organic material. In<br />

fact, mangrove root crabs may affect the<br />

distribution red mangroves. Little is known<br />

about the mangrove root crabs on Spanish<br />

Lookout Caye.<br />

Physical Description<br />

According to Kaplan (1984), the root crab is<br />

about 2.5 inches long and has stalked eyes, a<br />

brown carapace with white spots, and hairy red<br />

legs with white or yellow spots. Female crabs<br />

have a wide abdomen for egg storage, and males<br />

have a thin abdomen (Coulombe 1984). At<br />

sexual maturity, females and males are roughly<br />

the same size (Oliveira 2004).<br />

Conservation<br />

Any impact that significantly alters the red<br />

mangrove ecosystem may affect the red<br />

mangrove root crab because of its dependence<br />

on mangroves shelter and food. Worldwide,<br />

Figure 1. G. cruentata under red mangrove root on SLC<br />

mangroves are threatened by sea level rise,<br />

development, pollution, dredging, and disease.<br />

In Belize, the greatest threat is probably<br />

unsustainable development.<br />

Behavior<br />

When crabs mate, the female stores the male’s<br />

sperm until she releases her eggs. The eggs are<br />

kept in a mass (sponge) on the abdomen of the<br />

female and released simultaneously with the<br />

stored sperm, which flows over and fertilizes the<br />

eggs. Zygotes develop through four phases:<br />

fertilized egg, zoea larvae, megalops larvae, and<br />

adult (Coulombe 1984). Through personal<br />

observations we discovered that when mangrove<br />

root crabs are disturbed these crabs remain<br />

motionless for a short period of time before<br />

moving quickly sideways into a nearby burrow.<br />

If the crab does not automatically move away, it<br />

will bring its front claws forward in a defensive<br />

position.<br />

Figure 2. G. cruentata in Dominican Republic (photo by<br />

Pedro Genaro Rodriguez)<br />

Hugh Parkey Foundation for Marine Awareness & Education (belizeadventurelodge.com)<br />

<strong>Sirenian</strong> <strong>International</strong> (sirenian.org)


OUR INVESTIGATION<br />

Figure 2. G. crutentata on mud flat at SLC, 6 June 2009<br />

Introduction<br />

Red mangrove root crabs provide food for other<br />

species and reduce the insect population within<br />

the red mangrove ecosystem. By eating the leaf<br />

litter and organic material around the mangrove<br />

roots, they contribute to the recycling of<br />

nutrients. During our island orientation walk, we<br />

noticed these beautiful crabs and decided to<br />

investigate them further.<br />

Methods<br />

We conducted surveys each morning at 0600hrs<br />

for ten days during June 2009. For<br />

approximately one hour each morning, we<br />

walked the SLC central causeway, scanning the<br />

area on either side of the causeway, to observe<br />

the location and behavior of red mangrove root<br />

crabs.<br />

Results<br />

During low tides, G. crutentata were both near<br />

the mangrove roots and on the mud flat some<br />

distance away from the mangrove roots. At high<br />

tide, when there was no mud flat and the<br />

mangrove roots are partially submerged, crabs<br />

were observed only on structures above the<br />

water line, including roots, cement blocks,<br />

wooden planks, pimento stakes, and piles of<br />

emergent mud. They were predictably observed<br />

at certain locations and generally found alone or<br />

at least 30 cm away from any other crab.<br />

To better understand the distribution and<br />

behavior of G. crutentata on SLC, we propose<br />

the following sampling design for future<br />

students.<br />

Effect of Tidal Variation on Distribution and<br />

abundances of the Red Mangrove Root Crab<br />

on Spanish Lookout Caye, Belize<br />

Hypothesis: Tidal variations influence the<br />

abundance of the Red Mangrove Root Crab<br />

(Goniopsis cruentata) in the Spanish<br />

Lookout Cayes, Belize.<br />

Methods: Collect data during two complete<br />

lunar cycles, one during the dry season<br />

(February-April) and the other during the wet<br />

season (May-November). Conduct walking<br />

surveys three times a day at high tide, low tide,<br />

and intermediate tide. Each survey should<br />

include one transect along the central causeway<br />

in a single direction, one person on either side of<br />

the board walk scanning for crabs. Each person<br />

will observe both the area around the roots and<br />

in the water. The number and location of crabs<br />

will be recorded on a map of the area. At the<br />

beginning and end of each bridge along the<br />

causeway, water depth, salinity, temperature,<br />

and wind speed will be recorded. General<br />

observations, such as precipitation and debris<br />

will also be noted.<br />

If our hypothesis is true, then we expect more<br />

sightings of red mangrove root crabs at low tide<br />

than at high or intermediate tides.<br />

Literature Cited<br />

Coulombe, D. A. 1984. The Seaside Naturalist.<br />

Simon and Schuster. New York, New York. 139.<br />

Feller, I. C. 1996. Effects of nutrient enrichment<br />

on leaf anatomy of dwarf Rhizophora mangle<br />

L.(red mangrove). Biotropica 28: 13-22.<br />

Feller, I. C., Sitnik, M. 1996. Mangrove Ecology<br />

Workshop Manual.<br />

Kaplan, E. H. 1988. Southeastern and Caribbean<br />

Seashores. Houghton Mifflin Company. Boston,<br />

New York. Plate 22.<br />

Oliveira, N. F., Coelho, P. A. 2004. Maturidade<br />

sexual fisiológica em Goniopsis cruentata<br />

(Latreille) Crustacea, Brachyura, Grapsidae) no<br />

Estuário do Paripe, Pernambuco, Brasil. Revista<br />

Brasileira de Zoologia 21 (4): 1011<strong>–</strong>1015.<br />

Hugh Parkey Foundation for Marine Awareness & Education (belizeadventurelodge.com)<br />

<strong>Sirenian</strong> <strong>International</strong> (sirenian.org)


Cushion Sea Star<br />

Oreaster reticulates<br />

Jessica Barth and Aarin C. Allen<br />

Edited by Caryn Self-Sullivan, Ph.D.<br />

Ecology & Behavior of Manatees & Dolphins in<br />

Belize, Class of 2009 (contact: caryns@sirenian.org)<br />

Taxonomy<br />

Kingdom Animalia<br />

Phylum Echinodermata<br />

Class Asteroidea<br />

Order Valvatida<br />

Family Oreasteridae<br />

Genus Oreaster<br />

Species Oreaster reticulatus<br />

Ecology<br />

Cushion sea stars are found in the Atlantic<br />

waters from the Carolinas to Brazil (Puglisi,<br />

2008). Juveniles are green in color and can<br />

reach up 15cm in diameter (Kaplan 1982).<br />

Adults are tan, orange or rust in color and can<br />

grow to 50cm (Figure 1).<br />

O. reticulates is an omnivore, feeding on<br />

copepods, crab larvae and sponges (Puglisi<br />

2008). They are found in the calm shallow<br />

sandy flats around turtle grass, usually in<br />

shallow waters (Kaplan 1982). However, they<br />

have been found at depths of 37 meters (Puglisi<br />

2008).<br />

In Belize, they are ubiquitous in the seagrass<br />

beds where they are considered detritus feeders<br />

(C. Self-Sullivan, pers. comm.)<br />

Figure 1. Juvenile, adult, not to scale (Belize 2009).<br />

Behavior<br />

While feeding, cushion sea stars use their<br />

suction cup-like tube feet to wrap around their<br />

prey. In times when food availability is low the<br />

Figure 2. Six-armed specimen (Belize 2009).<br />

cushion sea star will prevent starvation by<br />

reabsorbing its own body tissue, which<br />

ultimately leads to a decrease in the sea stars<br />

overall size (Puglisi 2008).<br />

The species reproduces annually in the summer<br />

but can be found spawning continuously in the<br />

warm waters of the tropics (Puglisi 2008). The<br />

larvae can travel far distances in the ocean<br />

current before settling into a sea grass bed<br />

(Puglisi 2008).<br />

If a sea star is injured, it has the ability to<br />

regenerate its arms, as long as some portion of<br />

the central disk remains viable (Kaplan 1982),<br />

occasionally resulting in the production of an<br />

extra arm (Figure 2).<br />

Conservation<br />

O. reticulates was once the most common sea<br />

star in the Caribbean but is currently found most<br />

often in areas where the human population is<br />

relatively low. The species is considered to be<br />

rare in areas of high human populations due to<br />

being over harvested for souvenirs and<br />

aquariums (Puglisi 2008, Kaplan 1982).<br />

Hugh Parkey Foundation for Marine Awareness & Education (belizeadventurelodge.com)<br />

<strong>Sirenian</strong> <strong>International</strong> (sirenian.org)


OUR INVESTIGATION<br />

Introduction<br />

The Hugh Parkey Foundation for Marine<br />

Awareness and Education hosted our 2-week<br />

field course in June of 2009. At our base station<br />

on Spanish Lookout Caye (SLC, a 186 acre<br />

mangrove island just east of Belize City), we<br />

immediately noticed many bright orange sea<br />

stars in the seagrass beds just off the western<br />

shore of the caye. However, we didn’t see any<br />

off the eastern shore. This led us to ask why<br />

more sea stars were observed on the western<br />

than the eastern side of this pristine mangrove<br />

island.<br />

Methods<br />

The eastern shore of SLC faces windward,<br />

overlooking the Belize Barrier Reef and<br />

Caribbean Sea. This side of the caye is subject<br />

to more wind and wave action than the western<br />

shore, which overlooks the Belize mainland,<br />

about 10 miles west of SLC. To confirm our<br />

initial observation, we snorkeled along both<br />

sides of the caye once a day for ten days,<br />

counting cushion sea stars. Our snorkel<br />

transects were approximately 100m long and<br />

approximately 5m offshore.<br />

During our snorkels, we also noted the color,<br />

size, and behavioral state of each individual sea<br />

star, along with the depth and substrate at which<br />

they were found. We simulated turbulent<br />

conditions by placing individuals on their backs,<br />

and observing their behavior.<br />

Results<br />

During the surveys, we found no sea stars off the<br />

eastern side of the island, but many sea stars off<br />

the western side. On average, we observed 11<br />

sea stars per day off the western shore (Figure<br />

3). Smaller sea stars, presumably juveniles,<br />

were green in color. Medium, presumably subadults,<br />

were green with brown and orange<br />

patches. Large, presumably adults, were tan,<br />

orange, or rust in color. Sea stars were most<br />

often found on the sandy flats outside of the<br />

turtle grass (Thalassia testudinum) beds, in<br />

water depths of 1-3 meters. When turned over,<br />

most were holding onto some sort of shell or<br />

rock with their tube-feet.<br />

Number of Sea Stars<br />

Observed<br />

Sea Star Observations<br />

Figure 3. Counts off the western shore of SLC.<br />

The sea star’s body was somewhat pliable when<br />

picked up underwater; however the body<br />

became rigid if removed from the water. When<br />

picked up and held for a few minutes, they<br />

would attached to our<br />

hand using feet-like<br />

suction tubes.<br />

However, if we swam<br />

with one in our hand,<br />

it would become<br />

limp. When we<br />

inverted an individual<br />

to simulate conditions<br />

that may affect a sea star in a rough<br />

environment, it took approximately three<br />

minutes for the sea star to turn back over.<br />

Discussion<br />

With only ten days in the field, we were unable<br />

to follow-up on the results of our inductive,<br />

hypothesis generating survey. If given a chance<br />

to deductively test our hypothesis that “sea stars<br />

select the western shore to avoid more turbulent<br />

waters”, we would set up an experiment under<br />

laboratory conditions. Perhaps we could set up a<br />

large tank to simulate the 2 environments found<br />

around the caye. If a sea star placed in the more<br />

turbulent environment moved towards the<br />

calmer environment, then our hypothesis would<br />

be supported.<br />

Literature Cited<br />

Kaplan, Eugene H., 1982. A Field Guide to Coral<br />

Reefs Caribbean and Florida. New York: Houghton<br />

Mifflin. <strong>17</strong>6 pp.<br />

Puglisi, Melany P. 2008. Oreaster reticulatus<br />

http://www.sms.si.edu/IRLSpec/Oreaster_reticulatus.<br />

htm. Downloaded 06/01/2009.<br />

Hugh Parkey Foundation for Marine Awareness & Education (belizeadventurelodge.com)<br />

<strong>Sirenian</strong> <strong>International</strong> (sirenian.org)<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

1 2 3 4 5 6 7 8 9 10<br />

Day of Observation


Fiddler Crab<br />

Uca minax<br />

Whitney Montgomery & Daniel Hill<br />

Ecology & Behavior of Manatees & Dolphins in<br />

Belize, Class of 2010 (contact: caryns@sirenian.org)<br />

Taxonomy<br />

Kingdom Animalia<br />

Phylum Arthropoda<br />

Subphylum Crustacea<br />

Class Malacostraca<br />

Order Decapoda<br />

Family Ocypodidae<br />

Species Uca minax<br />

Ecology<br />

Fiddler crabs inhabit a wide range of tropical<br />

regions. In Belize, their habitats are mangroves<br />

and salt marshes. Fiddler crabs are extremely<br />

populous species. The semi-terrestrial species is<br />

diverse in color and habitat, with the most<br />

distinguishing characteristic being the males’<br />

asymmetrical claws (Fig. 1). They use the large<br />

claw to signal to other crabs within their<br />

vicinity, to attract a mate, and as a weapon used<br />

in combat between other males (Magnhagen,<br />

1991). These behaviors could draw attention to<br />

them from predators. The claw can be up to<br />

50% of an adult male’s body weight. If a claw is<br />

lost, the smaller claw of some species will grow<br />

larger and the crab will regenerate the lost claw<br />

with a new, smaller one.<br />

Figure 1. Fiddler crabs during experiment. (Belize, 2010)<br />

fiddler crab (Belize 2010)<br />

Figure<br />

2. Adult<br />

male<br />

Behavior<br />

Fiddler crabs tend to feed near their burrows, so<br />

that if they feel threatened, they will retreat into<br />

it to seek shelter. The crabs use their small<br />

claws to pick up pieces of sand, filtering out<br />

microscopic bits of algae, microbes, or fungus<br />

with their mouthparts. Crabs live about two<br />

years and molt their shells indeterminately (one<br />

to two times per year) as they continue to grow<br />

throughout their lives. Males wave their larger<br />

claw to attract female attention when ready to<br />

reproduce. Fiddler crab eggs are fertilized<br />

externally and the female fiddler crabs carry<br />

their eggs on their underside in a large mass.<br />

She keeps them safe in her burrow for a two<br />

week period and then journeys down to the edge<br />

of the sea to release the eggs into the receding<br />

tide. The larvae remain in a planktonic state for<br />

a further two weeks before venturing onto land<br />

and setting up a territory of their own.<br />

Conservation<br />

While fiddler crabs are not seriously threatened,<br />

their habitats are under pressure. As human<br />

populations continue to expand, they are<br />

destroying the mangroves and salt marshes<br />

where the crabs live. Fiddler crabs may serve as<br />

an important indicator species for these places.<br />

Fiddler crabs are important to the conservation<br />

of their ecosystems. Some experts believe that<br />

by continually sifting through the sand for food,<br />

their feeding habits aerate the substrate and<br />

prevent anaerobic conditions (Montague, D. L.<br />

1980).


OUR INVESTIGATION<br />

Introduction<br />

On a 13-day Field Biology class, the Hugh Parkey<br />

Foundation for Marine Awareness and Education<br />

was our host for the May 26 <strong>–</strong>June 8, 2010<br />

educational trip. Staying and studying on the<br />

Drowned Cayes (a mangrove island east of the<br />

Belize mainland), fiddler crabs are one of the easiest<br />

species to first recognize. After several days of<br />

observation of their behaviors and movements, the<br />

question of “Does the handedness of a fiddler crab<br />

affect its movement when threatened?” was<br />

developed. In other words, do crabs with a larger<br />

left claw tend to move towards the left side when<br />

threatened?<br />

Method<br />

After a tour of the island, we determined two prime<br />

spots for observation and collection of data. On the<br />

west side of the island, the mangroves and other<br />

vegetation provide protection for burrowing crabs<br />

with still access to sediments in which they eat. We<br />

studied the red-jointed fiddler crab (Uca minax) for a<br />

period of 2-h over the course of 2-d, with special<br />

emphasis on movement and behavior when<br />

threatened. We then collected a total of 20 fiddler<br />

crabs over another 2-d period (nleft claw=10 and nright<br />

claw=9). We used a ruler to determine the claw<br />

length of each crab (Fig. 2) and noted whether the<br />

larger claw was on the right or left (handedness).<br />

The crab was placed into a plastic container and<br />

allowed time for to calm down from the stress of<br />

human handling. We then “threatened” the crab by<br />

using the ruler to come at the crab facing its front<br />

(head) side. Using the same ruler, we measured the<br />

distance it traveled and noted the direction in which<br />

it moved. Ten of the fiddler crabs were collected<br />

from the West side of the island within an area and<br />

the other 10 were collected from a similar site on the<br />

East side of the island. They were taken from two<br />

sites to ensure we had a wide variety of crabs. The<br />

crabs were returned to their original sites upon<br />

completion of the experiment.<br />

Results<br />

Of the 20 crabs tested, 11 moved in the direction<br />

opposite their large claw (55%) and 9 moved in the<br />

direction of their large claw (45%). We found that<br />

the direction moved in relation to the larger claw<br />

was not significant. Using a chi-square test (df=1),<br />

we found crabs did not move significantly in the<br />

direction of their larger claw (Fig. 3, =0.2, p ><br />

0.05).<br />

<br />

<br />

<br />

Figure 3. The number of crabs and direction of crab movement s in<br />

relation to their larger claw.<br />

Discussion<br />

Our hypothesis that fiddler crabs would be more<br />

likely to move in the direction of their larger claw<br />

was not supported. The sample size was reasonable<br />

given the time constraint and if given the<br />

opportunity to redo the experiment, a larger sample<br />

size would allow for a stronger test of the<br />

hypothesis. While our findings showed no size<br />

sided preference in movement, a few observations<br />

suggested that fiddler crabs might move away from<br />

their large claw when threatened. Some research<br />

suggests that the large claw increases vulnerability<br />

to predation (Martin & Lopez, 1999), while other<br />

research suggests the claw might actually be used to<br />

ward off potential avian predators (Bildstein et all,<br />

1989). More research on the effect of fiddler crab<br />

movement while threatened could test this<br />

hypothesis.<br />

References<br />

Bildstein, K. L., McDowell, S. G. & Brisbin, I.L. 1989.<br />

Consequences of sexual dimorphism in sand fiddler<br />

crabs, Uca pugilator: differential vulnerability to<br />

avian predation. Animal Behavior, 37, 133-139.<br />

Magnhagen, C. 1991. Predation risk as a cost of<br />

reproduction. Trends in Ecology and Evolution, 6,<br />

183-186.<br />

Martin, J. & Lopez, P. 1999. When to come out from a<br />

refuge: risk-sensitive and state-dependent decisions<br />

in an alpine lizard. Behavioral Ecology, 10, 487-<br />

492.<br />

Montague, D. L. 1980. A natural history of temperate<br />

western Atlantic fiddler crabs (genus Uca) with<br />

reference to their impact on the salt marsh.<br />

Contributions in Marine Science, 23, 25-55.


Patch Reef Habitats<br />

Kelly Haun and Emy Rodriguez<br />

Edited by Caryn Self-Sullivan<br />

Ecology & Behavior of Manatees & Dolphins in<br />

Belize, Class of 2010 (contact: caryns@sirenian.org)<br />

Ecology<br />

Corals are defined as small animals that attach<br />

themselves on hard surfaces to the sea floor.<br />

They colonize in the thousands and build<br />

skeletal structures of calcium carbonate, which<br />

produces the limestone foundation of coral<br />

reefs (Humann & Deloach, 2002). Hard<br />

corals are reef building corals that provide<br />

shelter and food for juvenile fish.<br />

Several types of coral reefs exist, which are<br />

defined by their general structure. Patch reefs<br />

represent one type of coral zone found near<br />

shore and are commonly surrounded by<br />

seagrass, sand, or algae (Mumby, 1999).<br />

These reefs are composed of hard and soft<br />

corals and rock or rubble, which form the<br />

concrete foundation on the sea floor.<br />

Patch reefs provide protective niches<br />

and food sources for many species of<br />

vertebrates and invertebrates (Kaplan, 1982).<br />

Conservation<br />

Coral reefs are in a state of decline around the<br />

world due to anthropogenic impacts that<br />

include overfishing, increased runoff and<br />

pollution from land, rising concentrations of<br />

carbon dioxide and invasive species (Turnball<br />

& Harborne, 2000). Overgrowth of algae,<br />

sponges and tunicates appears to have<br />

increased at many locations in the Caribbean.<br />

The growing problem can be partially<br />

contributed to the excess of nitrogen and other<br />

nutrients carried into the sea by runoff which<br />

stimulates the growth of plankton. The<br />

increase of plankton provides food to such<br />

organisms, which compete for space on corals<br />

and may eventually cause corals to die<br />

(Humann & Deloach, 2002). No previous<br />

studies have been made to analyze the reef<br />

system around the Drowned Cayes or<br />

document how these impacts may be affecting<br />

the reef ecosystem.<br />

View of reef patch from the surface.<br />

Introduction<br />

Our Investigation<br />

The Belize Barrier Reef System fringes along<br />

the eastern side of the Spanish Lookout Caye,<br />

which is a mangrove island surrounded by<br />

seagrass beds. Patch reefs are intermingled<br />

throughout and contribute to this complex and<br />

intricate environment.<br />

With the help of our guide, Denroy Usher, we<br />

surveyed the coastline around the caye and<br />

found several patch reefs varying in depth<br />

from 4 ft to 18 ft. These reefs have not been<br />

previously studied and an understanding of<br />

their structure and role can help contribute to<br />

conservation efforts in the area.<br />

Our objectives are to draw up preliminary data<br />

on the percentage of substrate types found in a<br />

coral patch and determine the relative<br />

abundance of fish in relation to the distance<br />

from a coral patch. We predict that the<br />

number of fish will decrease the further we<br />

move from the patch.<br />

Hugh Parkey Foundation for Marine Awareness & Education (belizeadventurelodge.com)<br />

<strong>Sirenian</strong> <strong>International</strong> (sirenian.org)


Method<br />

We selected a patch reef located on the south<br />

west point of the caye approximately 150 ft<br />

from the coastline. The depth of the coral<br />

patch was 4.7 ft with 100% water clarity<br />

which enabled ease of data collection. We<br />

randomly selected a patch head and set up a<br />

10 ft x 10ft study site. We randomly sampled<br />

four quadrants within the site to determine the<br />

percentage of substrates within the patch. A<br />

line transect was positioned from the center of<br />

the site east and west along the contour of the<br />

coastline each 40 feet from the center, totaling<br />

80 feet across. Fish counts were conducted<br />

every 10ft radiating from the center of the<br />

study site.<br />

Results<br />

Our findings show that substrate cover varied<br />

between quadrants 1 and 2. After calculating<br />

the total percent cover of both quadrants we<br />

determined the reef patch substrate to be<br />

composed of 40% rock/rubble, 22.5% hard<br />

coral and 16.25% algae/sponge (See Table 1<br />

in appendix).Figure 1 shows the data obtained<br />

during the fish line transect, showing a general<br />

trend whereby the total number of fish<br />

decreased as the distance from the coral patch<br />

increased. However, the east transect shows a<br />

decreased number of fish at 10 ft from the<br />

patch reef.<br />

Figure 1. Number of fish located at 10 ft increments from<br />

the center of the research site.<br />

Discussion<br />

This study attempted to draw up preliminary<br />

data on the percentage of substrate types<br />

found in a coral patches. However, due to<br />

limited time constraints we were unable to set<br />

up a control site with which to provide a<br />

comparative analysis.<br />

Our fish counts confirm our initial hypothesis<br />

that the relative abundance of fish decreased<br />

in relation to distance from the coral patch.<br />

This demonstrates the importance of patch<br />

reefs as a shelter and as a food source for fish,<br />

and suggests fish do not migrate far from these<br />

rocky/rubble and hard coral substrates.<br />

We believe that the slight discrepancy in data<br />

on the east transect line (10 ft distance) is a<br />

result of disturbance caused while performing<br />

the fish counts, as snorkeling became more<br />

challenging at shallower depths.<br />

We observed algal growth on many hard<br />

corals as well as a small amount of disease<br />

which may be indicative of increased nitrogen<br />

levels, as well as other anthropogenic stresses,<br />

and should be further investigated.<br />

Literature Cited<br />

Humann, P., & Deloach, N. (2002). Reef Coral<br />

Identification. Jacksonville: New World Publications.<br />

Kaplan, E. H. (1982). Ecology of the Coral Reef. In<br />

E. H. Kaplan, A Field Guide to Coral Reefs,<br />

Carribean and Florida (pp. 101-120). New York:<br />

Houghton Mifflin Company.<br />

Mumby, P. J. (1999). Classification Scheme for<br />

Manatee Habitats of Belize. London: Centre for<br />

Tropical Coastal Management Studies.<br />

Turnball, C., & Harborne, A. (2000). Summary of<br />

Coral Cay Conservation's Atlantic and Gulf Rapid<br />

Reef Assessment Data From Turneff Atoll, Belize.<br />

London: Coral Cay Conservation LTD.<br />

Hugh Parkey Foundation for Marine Awareness & Education (belizeadventurelodge.com)<br />

<strong>Sirenian</strong> <strong>International</strong> (sirenian.org)


Method<br />

We selected a patch reef located on the south<br />

west point of the caye approximately 150 ft<br />

from the coastline. The depth of the coral<br />

patch was 4.7 ft with 100% water clarity<br />

which enabled ease of data collection. We<br />

randomly selected a patch head and set up a<br />

10 ft x 10ft study site. We randomly sampled<br />

four quadrants within the site to determine the<br />

percentage of substrates within the patch. A<br />

line transect was positioned from the center of<br />

the site east and west along the contour of the<br />

coastline each 40 feet from the center, totaling<br />

80 feet across. Fish counts were conducted<br />

every 10ft radiating from the center of the<br />

study site.<br />

Results<br />

Our findings show that substrate cover varied<br />

between quadrants 1 and 2. After calculating<br />

the total percent cover of both quadrants we<br />

determined the reef patch substrate to be<br />

composed of 40% rock/rubble, 22.5% hard<br />

coral and 16.25% algae/sponge (See Table 1<br />

in appendix).Figure 1 shows the data obtained<br />

during the fish line transect, showing a general<br />

trend whereby the total number of fish<br />

decreased as the distance from the coral patch<br />

increased. However, the east transect shows a<br />

decreased number of fish at 10 ft from the<br />

patch reef.<br />

Figure 1. Number of fish located at 10 ft increments from<br />

the center of the research site.<br />

Discussion<br />

This study attempted to draw up preliminary<br />

data on the percentage of substrate types<br />

found in a coral patches. However, due to<br />

limited time constraints we were unable to set<br />

up a control site with which to provide a<br />

comparative analysis.<br />

Our fish counts confirm our initial hypothesis<br />

that the relative abundance of fish decreased<br />

in relation to distance from the coral patch.<br />

This demonstrates the importance of patch<br />

reefs as a shelter and as a food source for fish,<br />

and suggests fish do not migrate far from these<br />

rocky/rubble and hard coral substrates.<br />

We believe that the slight discrepancy in data<br />

on the east transect line (10 ft distance) is a<br />

result of disturbance caused while performing<br />

the fish counts, as snorkeling became more<br />

challenging at shallower depths.<br />

We observed algal growth on many hard<br />

corals as well as a small amount of disease<br />

which may be indicative of increased nitrogen<br />

levels, as well as other anthropogenic stresses,<br />

and should be further investigated.<br />

Literature Cited<br />

Humann, P., & Deloach, N. (2002). Reef Coral<br />

Identification. Jacksonville: New World Publications.<br />

Kaplan, E. H. (1982). Ecology of the Coral Reef. In<br />

E. H. Kaplan, A Field Guide to Coral Reefs,<br />

Carribean and Florida (pp. 101-120). New York:<br />

Houghton Mifflin Company.<br />

Mumby, P. J. (1999). Classification Scheme for<br />

Manatee Habitats of Belize. London: Centre for<br />

Tropical Coastal Management Studies.<br />

Turnball, C., & Harborne, A. (2000). Summary of<br />

Coral Cay Conservation's Atlantic and Gulf Rapid<br />

Reef Assessment Data From Turneff Atoll, Belize.<br />

London: Coral Cay Conservation LTD.<br />

Hugh Parkey Foundation for Marine Awareness & Education (belizeadventurelodge.com)<br />

<strong>Sirenian</strong> <strong>International</strong> (sirenian.org)


139<br />

140<br />

141<br />

142<br />

143<br />

144<br />

145<br />

146<br />

147<br />

148<br />

149<br />

Average BML Average AML


Variegated Sea Urchin<br />

Lytechinus variegatus<br />

Rachelle Boucher, Elizabeth Ferrell, and<br />

Samantha Egelhoff<br />

Edited by Caryn Self-Sullivan, Ph.D<br />

& Dr. Bruce Schulte, Ph. D<br />

Ecology & Behavior of Manatees & Dolphins in<br />

Belize, Class of 2010 (contact:<br />

caryns@sirenian.org)<br />

Taxonomy<br />

Kingdom: Animalia<br />

Phylum: Echinodermata<br />

Class: Echinoidea<br />

Order: Temnopleuroida<br />

Family: Toxopneustidae<br />

Genus: Lytechinus<br />

Species: Lytechinus variegatus<br />

Ecology<br />

Commonly known as the green sea urchin,<br />

Lytechinus variegatus is found in waters from<br />

Bermuda to Brazil, including the Caribbean Sea<br />

(Norris 2003). In their adult form, they are<br />

mainly white to green in color, although some<br />

have a pinkish-purple hue. They are covered in<br />

short spines that can regenerate. While their size<br />

usually varies from 1-12 cm, some can grow as<br />

large as 36 cm (Barnes 1982).<br />

Purple Variation of L. variegatus<br />

L. variegatus is normally found in calm, shallow<br />

waters, although they can be found in waters up<br />

to 50m deep (Norris 2003). It feeds on seagrass,<br />

mainly Thalassia sp., by using its tube feet and<br />

Aristotle’s Lantern (Norris 2003). The tube feet<br />

act as an anchor and hold the green sea urchin in<br />

White variation of L. variegatus<br />

place, while the mouth of the sea urchin, known<br />

as Aristotle’s Lantern, which is composed of<br />

strong joints and teeth, can scrape its food off of<br />

rocks or other hard surfaces (Barnes 1982).<br />

Behavior<br />

The green sea urchin reproduces by releasing<br />

unfertilized eggs and sperm into its water column,<br />

where the eggs are fertilized and develop into<br />

larvae (Norris 2003). The larvae undergo a<br />

complex metamorphosis to reach the adult form.<br />

L. variegatus use their spines and tube feet to<br />

cover their exterior with surrounding debris,<br />

serving as protection from predators and the sun’s<br />

rays.<br />

Conservation<br />

In Caribbean reef ecosystems, sea urchins have<br />

the capability of population explosion if one of<br />

their key predators is taken out of the system.<br />

Since their main food source is turtle grass<br />

(Thalassia sp.), a large urchin population can<br />

easily clear out whole patches of seagrass on<br />

which other animals, such as manatees and sea<br />

turtles, depend. However if sea urchins are taken<br />

out of a reef ecosystem, the seagrass patches that<br />

they normally keep under control can grow<br />

exponentially in some reef regions and kill off<br />

fragile corals. Jackson, however, mentions that<br />

Diadema urchins have always been abundant in<br />

their environment since pre-colonization, and<br />

predator control might not even be a factor<br />

(1997).


OUR INVESTIGATION<br />

Introduction<br />

We found L. variegatus to be quite common<br />

along the sea wall on the west side of Spanish<br />

Lookout Caye off the coast of Belize City,<br />

Belize. While observing them, we noticed that<br />

they were always hidden underneath rocks or<br />

covered with seagrass or shells, perhaps done in<br />

order to provide protection against predators or<br />

to protect them from the sun or other physical<br />

factors (Verling, 2004). The objective of this<br />

research was to see if the size of the sea urchin<br />

determined how fast the urchin recovered itself<br />

with debris, or if the size was related to how far<br />

they traveled away from their initial spot once<br />

uncovered. Smaller sea urchins are in need of<br />

more protection because they are more<br />

susceptible to threats. Therefore, we expected<br />

them to travel further or in less time in order to<br />

find protection.<br />

Methods<br />

Over a four-day period, we noted abiotic<br />

conditions and searched for different sized sea<br />

urchins along the shallow sandy and rocky<br />

bottoms. We removed all debris on 10 urchins<br />

using tweezers and tested 3 additional control<br />

urchins (n=13). Two controls were untouched,<br />

and one was touched but no debris was removed.<br />

Every two minutes for twenty minutes we<br />

measured the distance each traveled from its<br />

initial position. After observing for 20 minutes,<br />

we measured the diameter of each urchin, not<br />

including spines, and noted its coloration. Two<br />

graphs were plotted to see if the size of the<br />

urchins had an effect on the time of movement<br />

or distance traveled in order to recover<br />

themselves with debris.<br />

Results<br />

We found that there was no significant<br />

relationship between size and total distance<br />

(F(1,8)=1.32, P=0.28) (Fig 1) or size and time of<br />

movement (F(1,8)=0.38, P=0.55) (Fig 2). We<br />

tested a variety of juvenile-sized urchins, but<br />

each urchin had its own individual response. The<br />

control urchins suggested that there was no<br />

relationship between human interaction and<br />

individual urchin movement, only that they<br />

sought protection. All urchins either recovered<br />

themselves with debris or moved under rocks in<br />

place of coverings within the 20-minute time<br />

frame after we removed their initial debris.<br />

Figure 1. Total Distance Moved vs. Size<br />

Figure 2. Total Time of Movement vs. Size<br />

Discussion<br />

Time and distance traveled by each individual<br />

urchin appeared to be to dependent on how<br />

quickly they were able to recover themselves,<br />

whether with debris or under a rock; not their size.<br />

Further research should be done using a larger<br />

sample size including adult urchins in varying<br />

habitats. Since most of the urchins were juveniles<br />

and found in shallow water, this might have<br />

skewed our data and conclusions.<br />

Literature<br />

Barnes, Robert D. (1982). Invertebrate Zoology.<br />

Philadelphia, PA: Holt-Saunders <strong>International</strong>.<br />

Pp 961-981. ISBN 0-03-056747-5.<br />

Jackson, J.B.C. (1997). Reefs since Columbus. Coral<br />

Reefs: S23-S32.<br />

Norris, Amy. (2003). Green Sea Urchin Lytechinus<br />

variegatus. Marine Invertebrates of Bermuda:<br />

1-7.<br />

Verling, E. D.K.A. (2004). The dynamics of covering<br />

behavior in dominant Echinoid populations<br />

from American and European West coasts.<br />

Marine Biology: 191-206.


Red Fin Needlefish<br />

Strongylura natata<br />

Jamie Hennis and Carol Lacey<br />

Edited by Bruce A. Schulte, Ph.D. and<br />

Caryn Self-Sullivan, Ph.D.<br />

Ecology & Behavior of Manatees &<br />

Dolphins in Belize, Class of 2010.<br />

(contact: caryns@sirenian.org)<br />

Taxonomy<br />

Kingdom Animalia<br />

Phylum Chordata<br />

Super Class Osteichthyes<br />

Class Actinopterygii<br />

Order Beloniformes<br />

Family Belonidae<br />

Genus Strongylura<br />

Species Strongylura natata<br />

Ecology<br />

The redfin needlefish is a marine fish<br />

located in warm, clear, coastal waters<br />

(Humann 2002). They can be seen<br />

solitary or in small groups in mangroves<br />

and sea grass ecosystems (Sweat 2009).<br />

Redfin needlefish are about 9 times their<br />

length than their width with an elongated<br />

needle-like bill. All species of needlefish<br />

are blue-green in color dorsally with<br />

silver sides. The redfin needlefish gets<br />

its name because the caudal and anal fins<br />

appear red in color. They are similar to<br />

other species in this family because their<br />

upper jaw is slightly shorter then the<br />

lower jaw, they have a short tail and the<br />

lower lobe is slightly longer than the<br />

upper lobe. We could find no<br />

information regarding conservation<br />

issues of needlefish.<br />

Figure 1. Needlefish at rest (photo by Carol<br />

Lacey).<br />

Behavior<br />

Needlefish are considered surface feeders.<br />

They are generally found just below the<br />

water surface where feeding is convenient for<br />

them due to their lateral line being located<br />

ventrally on their body (Lovejoy 2000). They<br />

are known as ram feeders. They capture their<br />

prey by striking them with their beak and<br />

pinning them in their jaws. The organism is<br />

then released and positioned properly for<br />

consumption in a head first position. Redfin<br />

needlefish have low predation risk because of<br />

their blue-green counter shading with the<br />

water, camouflaging them from overhead<br />

birds. Their nicknamed the skipper because<br />

when frightened, they tend to leap out of the<br />

water and skip across the surface.<br />

Figure 2. Needlefish foraging (photo by Carol<br />

Lacey).


OUR RESEARCH FINDINGS<br />

Introduction<br />

The Hugh Parkey Foundation for Marine<br />

Awareness and Education hosted our 13day<br />

field course (May 26-June 8, 2010).<br />

Our base station was located at Spanish<br />

Lookout Caye in the Drowned Cayes<br />

east of Belize City. This is a mangrove<br />

island made up of 186 acres. We found a<br />

population of redfin needlefish near the<br />

boat dock. We observed these fish for<br />

two days prior to collecting data and<br />

noticed several types of behavior. These<br />

observations suggested that activity<br />

levels depend on the time of day. Our<br />

objective was to determine if the<br />

occurrence of redfin needlefish<br />

behaviors varied with time of day.<br />

Method<br />

The sample area was 152.7 sq ft between<br />

the restaurant and the manatee museum.<br />

This area is full of marine life offering<br />

food and calm water (except waves from<br />

boats docking). During our observations<br />

we identified three behaviors as follows:<br />

1) Resting - needlefish oriented to the<br />

current in a head first position (Fig. 1).<br />

2) Foraging - needlefish swim in a<br />

random, non-organized manner looking<br />

for food (Fig. 2). 3) Socialization -<br />

needlefish appear to be aware of others<br />

in the population and interactions or<br />

non-aggressive swim takes place.<br />

We observed needlefish three times each<br />

day: morning 6:30-8 am, afternoon 12-4<br />

pm, and evening 6-9 pm. We scanned<br />

fish behavior once per minute for fifteen<br />

minutes, only including calculations in<br />

our results when fish were present. The<br />

abiotic factors of tidal depth, current<br />

speed, and weather conditions also were<br />

recorded.<br />

Results<br />

Needlefish were found to be most socially<br />

active in the morning and afternoon, with the<br />

three behaviors occurring equally at night<br />

(Fig. 3). The only abiotic factor appearing to<br />

have an effect on needlefish behavior was<br />

weather; needlefish were absent during rains.<br />

Figure 3. Number of fish resting, foraging, and<br />

socializing three times during the day.<br />

Discussion<br />

We found that fish numbers did not differ<br />

with time of day, but there was a difference<br />

in behaviors. Behaviors were equal in<br />

occurrence at night with mornings showing<br />

the highest socializing. With further research<br />

we could develop a better understanding of<br />

the effect time of day and other abiotic<br />

factors have on needlefish behavior and<br />

activity.<br />

Literature Cited<br />

Humann, Paul. 2002, Reef Fish Identification<br />

Florida Caribbean Bahamas, pg. 58.<br />

Lovejoy, Nathan R., 2000. Systematics of<br />

Needlefishes and their Allies (Teleostei:<br />

Beloniformes). Volume 54:pg 1349-1362.<br />

Sweat, L.H. 2009, Redfin Needlefishes.<br />

Smithsonian Marine Station at Fort Pierce.<br />

Downloaded May 20, 2010.<br />

www.sms.si.edu/irlspec/strong_hotata.htm.


Fiddler Crabs and Hermit Crabs<br />

Uca minax and Pagurus annulipes<br />

Rachel Calhoun and Molly Martin<br />

Ecology & Behavior of Manatees & Dolphins in Belize,<br />

Class of 2010 (contact:cselfsullivan@sirenian.org)<br />

Taxonomy<br />

Kingodom Animalia<br />

Phylum Arthropoda<br />

Subphylum Crustacea<br />

Class Malacostraca<br />

Order Decapoda<br />

Family Ocypodidae<br />

Species Uca minax<br />

Superfamily Paguroidea<br />

Species Pagurus annulipes<br />

Ecology<br />

Fiddler crabs are found in mangroves, salt marshes,<br />

and on sandy or muddy beaches in the Southeastern<br />

U.S., the Gulf of Mexico, and the Carribean (Kaplan<br />

1988). Their distinctively asymmetric claws easily<br />

identify them. U. minax is a detritivore, consuming<br />

algae, microbes, fungus, and other decaying materials.<br />

They appear to be a common resident along the<br />

brackish intertidal mud flats, lagoons, and swamps of<br />

Belize.<br />

Similarly, terrestrial hermit crabs are found in tropical<br />

areas within the intertidal zone. They also are<br />

detritivores, feed primarily on algae and decaying<br />

materials. They lack a carapace, or shell, so they<br />

“borrow” one from snails, such as periwinkles or<br />

oyster drills. As hermit crabs grow in size, they have<br />

to locate a larger shell, abandoning the previous one.<br />

Aside from this, their soft-coiled abdomens, two pairs<br />

of walking legs, and their asymmetric claws, are also<br />

used to identify them.<br />

Behavior<br />

Fiddler crabs create small burrows in the sediment<br />

that are used for mating, sleeping, refuge from<br />

predators, and ‘hibernating’ during the winter. They<br />

are very active during the day, foraging for food and<br />

digging burrows; they will return to their burrows at<br />

night and during high tide, plugging the entrance with<br />

mud or sand. Shells are an important resource for<br />

hermit crabs, but empty shells are generally low<br />

supply in nature (Kellogg 1976) and may be a limited<br />

resource for hermit crabs (Turra and Denadai 2004).<br />

As a result, hermit crabs will often compete with one<br />

another for shells. Competition may occur in two<br />

Figure 1. Hermit crab on the beach.<br />

ways: 1) hermit crabs may use chemical cues to locate<br />

newly available shells (Mesce 1982),or 2) they may<br />

display ritualized agonistic shell fighting behaviors<br />

and subordinate other individuals of the same species<br />

(Hazlett 1966). This is known as interference.<br />

However, many hermit crab species coexist in coastal<br />

areas and may cause various degrees of niche overlap<br />

(Turra and Denadai 2001).<br />

Introduction<br />

Our research took place during our 13-day research<br />

trip as part of a field marine biology class. We stayed<br />

on a mangrove island in the Drowned Cayes east of<br />

the Belize mainland with the Hugh Parkey Foundation<br />

for Marine Awareness and Education. Both hermit<br />

crabs and fiddler crabs are abundant on the island and<br />

are easily found. After much observation around the<br />

island of both crab species we asked the question of<br />

“How do fiddler and hermit crabs abundances change<br />

with the tides?”<br />

Figure 2. Fiddler crab near a burrow.<br />

Methods<br />

We chose two different areas to use for our research<br />

according to where we found the most of each crab<br />

species. Both areas were counted twice a day for three<br />

days. The area for the fiddler crabs was in a marshy<br />

Hugh Parkey Foundation for Marine Awareness & Education (belizeadventruelodge.com)<br />

<strong>Sirenian</strong> <strong>International</strong> (sirenian.org)


area on the north part of the island near the lagoon.<br />

The area for the hermit crabs was on the east part of<br />

the island only a few meters from the edge of the<br />

beach. Presence or absence of a breeze, temperature<br />

estimation, precipitation, the substrate moisture, and<br />

tide were taken before each data collection. We noted<br />

the presence and activity levels of each species in and<br />

around the study area before and after collecting data.<br />

We counted the number of fiddler crab burrows within<br />

a 10x10 ft area within the marshy area. This area had<br />

a small mangrove tree that marked one corner.<br />

Vegetation coverage in the area was estimated by<br />

looking at what percentage of the soil was covered by<br />

vegetation. The study area was divided in half and<br />

each person counted the number of burrows on each<br />

side to make counting easier and switched which side<br />

they counted in the afternoon.<br />

In the second area where hermit crabs were counted<br />

we demarcated a 15x15 ft study area with a palm tree<br />

in the eastern corner of the box and the number of<br />

hermit crabs on the beach was recorded. Next, the<br />

number of hermit crabs either climbing the tree or<br />

hidden in the braches of the tree were counted. Hermit<br />

crabs in the tree often crowded in crevices and would<br />

move when we got near making them more difficult to<br />

count. Therefore, each person counted and then<br />

shared numbers and an average number between<br />

counts was agreed upon. Also observations of hermit<br />

crabs in the area especially those near the water of the<br />

beach were recorded.<br />

Results<br />

The density of both crab species did not vary much<br />

with tide level (Fig. 3). The high tide for the fiddlers<br />

had the lowest density but it was also the only data<br />

taken during high tide whereas three data sets were<br />

taken during low tide and two taken during an ebb<br />

tide.<br />

The average density of fiddler crabs per square foot<br />

was 13 times greater than the average density of<br />

hermit crabs per square foot although the fiddler crab<br />

area was smaller. Most of the hermit crabs found were<br />

located in the crevices of the palm tree in the area and<br />

only a total of eight hermit crabs were found on the<br />

beach over all three days. The fiddler crab burrows in<br />

the designated area averaged 128.7 for each data<br />

collection and hermits found in the palm tree averaged<br />

21 for each data collection.<br />

Discussion<br />

We did not observe any major differences in the<br />

abundance of either crab species during different<br />

times of the tide. We did see six times more fiddler<br />

crabs than hermit crabs in a smaller area but neither<br />

species fluctuated too much with the tides. It is hard to<br />

draw any absolute conclusions from our data because<br />

we only took six data sets over three days and did not<br />

cover all the tides fairly. If we had more time to take<br />

more sets we would have more reliable data.<br />

Additionally, we did not have enough data to do any<br />

statistical analyses, so no certain conclusions can be<br />

made. However, this study seems to suggest that tides<br />

do not seem to directly affect either species’ numbers<br />

or their activity. Interestingly, the high tide did not<br />

reach the study site of the hermit crabs, which means<br />

that their decrease in density or relocation to the tree<br />

does not appear to be directly correlated with tides.<br />

There appears to be another reason as to why hermit<br />

crabs hide in trees, which could be predator<br />

avoidance.<br />

References<br />

Hazlett, B.A., 1966. Social behavior of the Paguridae and<br />

Diogenidae of Curacao. Stud. Fauna Curacao Other<br />

Caribb. Isl. Vol. 23, 1 <strong>–</strong>143.<br />

Kaplan, E.H. 1988. “Fiddler Crabs of the Southeastern<br />

U.S., Gulf of Mexico, and Carribean.” Southeastern<br />

and Caribbean Seashores. p 334-338<br />

Kellogg, C.W., 1976. Gastropod shells: a potentially<br />

limiting resource for hermit crabs. J. Exp. Mar. Biol.<br />

Ecol.Vol. 22, 101<strong>–</strong>111.<br />

Mesce, K.A., 1982. Calcium-bearing objects elicit shell<br />

selection behavior in a hermit crab. Science. Vol. 215,<br />

993<strong>–</strong>995.<br />

Turra, A., Denadai, M.R., 2001. Desiccation tolerance of<br />

four sympatric tropical intertidal hermit crabs<br />

(Decapoda Anomura). Mar. Freshw. Behav. Physiol.<br />

Vol. 34, 227<strong>–</strong> 238.<br />

Turra, A., Denadai, M.R., 2004. Interference and<br />

exploitation components in interespecific competition<br />

between sympatric intertidal hermit crabs. Journal of<br />

Experimental Marine Biology and Ecology Vol. 310,<br />

183<strong>–</strong> 193.<br />

<br />

Hugh Parkey Foundation for Marine Awareness & Education (belizeadventruelodge.com)<br />

<strong>Sirenian</strong> <strong>International</strong> (sirenian.org)


Cushion Sea Star<br />

Oreaster reticulates<br />

Whitney Allen and Rory McGonigle<br />

Edited by: Dr. Caryn Self-Sullivan, Ph.D.<br />

& Dr. Bruce Schulte, Ph.D.<br />

Ecology & Behavior of Manatees & Dolphins in<br />

Belize, Class of 2010 (contact: cselfsullivan@sirenian.org)<br />

Taxonomy<br />

Kingdom: Animalia<br />

Phylum: Echinodermata<br />

Class: Asteroidea<br />

Order: Valvatida<br />

Family: Oreasteridae<br />

Genus: Oreaster<br />

Species: Oreaster reticulates<br />

Ecology<br />

Cushion sea stars are found to be fairly common<br />

in Belize and other parts of the Caribbean. They<br />

are usually located in shallow waters inhabited<br />

with sea grasses and/or sand (Hummann and<br />

Deloach, 2002). Sea stars are normally fivearmed<br />

creatures varying in colors from green<br />

and brown to dark red and orange (Fig. 1).<br />

Juveniles tend to be green with spots of brown<br />

developing on them as they get older. Full adults<br />

may be yellow, orange, or red, normally mixing<br />

at least two of those together.<br />

Figure 1. Mid-range Adult Sea Star (Belize 2010)<br />

Some sea stars have six legs, which is almost<br />

always a cause of one previously being torn off.<br />

<br />

When growing back, it grows back mutated and<br />

forms two arms in the place of one (D. Usher,<br />

field assistant, pers. comm.).<br />

Behavior<br />

Sea stars have suction-like tube feet that allow<br />

them to stay on the sea floor (Fig. 2). Their<br />

connective tissue allows them “to change in<br />

response to passive mechanical properties” that<br />

can be caused by numerous environmental<br />

factors (Motokawa, 1984). When picked up or<br />

otherwise displaced, the tube feet close and are<br />

not opened until after the initial threat subsides.<br />

The center of their body is the central location of<br />

the digestive system, where all food is brought<br />

up into the body. The tube feet also may pick up<br />

food particles; those particles are then<br />

transferred down the arm to the mouth at the<br />

body’s center. Sea stars are not rapidly moving<br />

organisms. Our evidence suggests that they<br />

experience site fidelity, meaning they can be<br />

found in the same or near the same location on a<br />

day-to-day basis. However, their location can be<br />

affected by the water’s turbulence and weather<br />

conditions.<br />

Figure 2. Six-Armed Sea Star (Belize 2010)<br />

Conservation<br />

Sea stars, with their normally shallow location<br />

and non-movement along the sea floor, are<br />

typically taken for granted. With no visible<br />

breathing, many people forget they are even a<br />

living creature. Therefore, many sea stars lose<br />

their life when brought above water for too long,<br />

and from there, some are even kept for souvenirs<br />

and decorations.


Introduction<br />

Our Investigation<br />

Before our arrival to Hugh Parkey’s Adventure<br />

Lodge, we were asked to do pre-experimental<br />

research of information relating to a topic shared<br />

with a fellow student. After actual research<br />

began, we started by observing certain locations<br />

of sea stars along the western shore of the island.<br />

However, there did not appear to be a preference<br />

of their dwelling, so we eliminated that aspect<br />

out of our research. We noted that each time we<br />

picked up a sea star off the sea floor, it closed up<br />

its tube feet. It appeared that the time it took to<br />

completely close up the tube feet varied from<br />

one individual to another. This observation lead<br />

us to the main question of our project: Does the<br />

size of a sea star affect the amount of time it<br />

takes for it to close up its tube feet?<br />

Methods<br />

All of the data collected was done via snorkeling<br />

along the western shore (See appendix). Once a<br />

sea star was spotted, it was removed from the<br />

sea floor, and at the point of first contact the<br />

stopwatch was started. The star was flipped over<br />

in one’s hand and the time was recorded after<br />

complete closure. The sea star was measured<br />

from the end of one arm to the very center of the<br />

body. It was then measured again from the<br />

center of the body to a different leg to obtain a<br />

general diameter (two radii) for the organism.<br />

The following were also noted: location (sea<br />

grass or sand), color, number of legs and eating<br />

behavior at the time. After the data was<br />

collected, the sea star was gently placed back in<br />

its original position on the sea floor.<br />

Results<br />

No juveniles could be found on the western side<br />

of the island, but there was still a 16cm-27.5cm<br />

range of diameters in the sample pool (Figure 3).<br />

The average diameter was 24.0 cm (± 3.07).<br />

Four seconds was the most common closing<br />

time, consisting of 25% of our data. The fastest<br />

closing time was two seconds, which was a star<br />

measuring 25.5 cm in diameter. A 28.5 cm star<br />

closed the slowest, taking a total of nine<br />

<br />

seconds. Our evidence shows a weak but<br />

significant relationship between the size of a sea<br />

star and the time it takes to close (r 2 =0.20, df= 1,<br />

23, F=5.63, P=0.026; raw data located in<br />

appendix).<br />

Figure 3. Size of sea star and the time to completely close its tube<br />

feet when picked up and inverted underwater.<br />

Discussion<br />

Our hypothesis (“The bigger the sea star is the<br />

quicker it will close up its tube feet.”) was not<br />

supported by our data. We expected a negative<br />

correlation, while our results produced one that<br />

is positive. However, our acquired R square<br />

value shows that that the time it takes a sea star<br />

to close is somewhat size dependent, with larger<br />

stars taking longer. The rate at which it protects<br />

itself could also be caused by either the<br />

experience of turbulent waters or even predation.<br />

A follow-up analysis could be done to continue<br />

to further test our hypothesis. If each sea star in<br />

our sample pool had been in the exact same<br />

location or raised under the exact same<br />

conditions, the role of size in its closure speed<br />

would be easier to isolate.<br />

References<br />

Humann, P. and N. Deloach. (2002). Reef<br />

Creature Identification, Sea Stars, p.<br />

366. New World Publications.<br />

Motokawa, T. (1984). Connective tissue catch in<br />

echinoderms. Biol. Rev. 59, 255-270.


Cushion Sea Star<br />

Oreaster reticulatus<br />

Blakely Rice<br />

Edited by: Caryn Self-Sullivan, Ph.D.<br />

Ecology and Behavior of Manatees and<br />

Dolphins in Belize, Class of 2011<br />

(contact: cselfsullivan@sirenian.org)<br />

Taxonomy<br />

Kingdom Animalia<br />

Phylum Echinodermata<br />

Class Asteroidea<br />

Order Valvatidas<br />

Family Oreasteridae<br />

Genus Oreaster<br />

Species Oreaster reticulatus<br />

Ecology<br />

The Cushion Sea Star (Oreaster<br />

reticulatus) is located throughout the<br />

Caribbean and are commonly found in<br />

shallow water seagrass and sand<br />

habitats (Scheibling, 1980). Adult sizes<br />

can range from 5 inches to 20 inches<br />

and juveniles from about 2 inches to 6<br />

inches (Kaplan, 1982; Rice, personal<br />

measurements). Coloration in adults is<br />

generally orange, yellow and red or any<br />

variation of the above. Juvenile<br />

coloration is a broad spectrum of greens<br />

(Figure 1).<br />

Oreaster is predominately a omnivorous<br />

grazer eating a variety of organisms<br />

including sponges and algae<br />

(Scheibling, 1982).<br />

Behavior<br />

Oreaster reticulates feeds by inserting<br />

its stomach into or onto its prey and<br />

secrets enzymes that digest the good,<br />

which is then absorbed (Kaplan, 1982).<br />

Their tube feet help them hold onto their<br />

prey. Oreasters many tube feet enable<br />

them to move quite quickly and cover<br />

significant distances in a short amount<br />

of time.<br />

Because sea stars reproduce asexually<br />

in a marine environment gametes meet<br />

mostly by chance (Kaplan, 1982). To<br />

increase chances of fertilization sea<br />

stars will release their gametes if other<br />

gametes are detected in the area<br />

(Kaplan, 1982).<br />

Conservation<br />

Even though Oreaster reticulatus is one<br />

of the most common sea stars in the<br />

Caribbean populations have become<br />

rare in some areas mainly due to over<br />

harvesting for the souvenir trade<br />

(Kaplan, 1982).


Introduction<br />

Sea stars have long been a favorite<br />

sight for snorkelers and divers, mainly<br />

because they are one of the only<br />

animals that does not bite or sting when<br />

handled. There only defensive behavior<br />

when being picked up is to become<br />

rigid, however after some time they will<br />

relax. This behavior inspired the<br />

question for my independent research<br />

project, how long does it take for a sea<br />

star to relax after becoming rigid and is<br />

there a difference between adults and<br />

juveniles. I hypothesized that juveniles<br />

will take more time to relax and the<br />

adults do.<br />

Methods<br />

I chose the western side of Spanish<br />

Lookout Caye (SLC) as my study<br />

location because of the persistence of<br />

sea stars in that area. To investigate my<br />

hypothesis I snorkeled along the sea<br />

wall and sea grass beds until I found a<br />

sea star. When a sea star was found it<br />

was picked up and turned over until all<br />

the tube feet had retracted and the<br />

animal was rigid. This was done to<br />

insure that the timing was accurate as it<br />

was found that some animals took<br />

longer to become rigid than others.<br />

Once fully rigid the animal was placed in<br />

my right, once it was on my hand timing<br />

began. Timing stopped when the first<br />

tube feet were present. The time was<br />

recorded on a slate and the animal was<br />

replaced in its original location, habitat<br />

location and size of the animal were also<br />

recorded.<br />

Results<br />

At the end of the study a total of 20<br />

adults and 6 juveniles were surveyed. A<br />

two-sample t-test for independent data<br />

was conducted to compare the latency<br />

(time) in adults and juveniles. There was<br />

a significant difference in the latency of<br />

adults and juveniles (t=2.34, df= 24,<br />

p=0.014). A one-way ANOVA was also<br />

conducted for post-HOC analysis<br />

regarding the latency with habitat type<br />

and latency with size. In regards to<br />

latency and habitat there was no<br />

significant difference in adults (f=0.45,<br />

df=18, p=0.6455) or juveniles (f=1.19,<br />

df= 5, p=0.3366). When looking at<br />

latency in regards to size there was a<br />

significant difference in adults (f=11.86,<br />

df=19, p=0.000596) but not a significant<br />

difference in juveniles (f=0.47,<br />

df=5,p=0.6644).<br />

Discussion<br />

My hypothesis that latency in juveniles<br />

would be longer than in adults was<br />

disproved. I found that latency in adults<br />

was actually longer than in juveniles.<br />

Continued studies need to be conducted<br />

to determine the reason for this<br />

behavior.<br />

Resources<br />

Kaplan, Eugene H., 1982. A field guide<br />

to Coral Reef. New York: Houghton<br />

Mifflin. P.<strong>17</strong>1-<strong>17</strong>6.<br />

Scheibling, R.E., 1982. Feeding habits<br />

of Oreaster reticulatus<br />

(Echinodermata: Asteroidea). Bulletin<br />

of Marine Sciences, 32, 504-510.<br />

Scheibling, R.E., 1980. Abundance,<br />

Spatial distribution and size structure<br />

of populations of Oreaster reticulatus<br />

(Echinodermata: Asterodiea) on Sand<br />

Bottoms. Marine Biology, 107-119.


Magnificent Feather-Duster<br />

Sabellastarte magnifica<br />

Sean Herbert<br />

Ecology, Behavior and Conservation of<br />

Manatees & Dolphins in Belize, Class of 2011<br />

(contact: caryns@sirenian.org)<br />

Taxonomy<br />

Kingdom Animalia<br />

Phylum Annelida<br />

Class Polychaeta<br />

Order Sabellida<br />

Family Sabellidae<br />

Genus Sabellastarte<br />

Species S. magnifica<br />

Ecology<br />

Feather duster worms inhabit reefs, mangroves,<br />

rocks, wrecks, and sand or gravel bottoms. They<br />

often grow from coral heads (Humann & DeLoach<br />

2002), and in Belize, the magnificent variety is often<br />

found attached to mangrove roots. Whatever the<br />

substrate, they are usually found sharing the space<br />

with various sessile animals such as sponges,<br />

mollusks and tunicates. S. magnifica are among the<br />

largest and most common feather dusters found in<br />

Belize. Those who may cringe at the word ‘worm’<br />

need not fear; their bodies remain hidden, encased in<br />

a parchment-like tube attached to the substrate. The<br />

worms construct their tubes using fine sand and<br />

mucus secreted from glands just below their head.<br />

The only visible portion of the worm is the crown of<br />

feather-like appendages, which can be used to easily<br />

identify magnifica because of its exclusive doublecrown<br />

arrangement. The ‘feathers’ are usually<br />

banded and come in a variety of colors, including<br />

brown, tan, gold, reddish purple and white (Humann<br />

& DeLoach 2002).<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Behavior<br />

Feather duster worms use their ‘feathers’, known<br />

as radioles, as both gills and filters to capture<br />

plankton <strong>–</strong> their cuisine of choice. Trapped<br />

microscopic food is brought to its mouth at the<br />

center of its crown of ‘feathers’ (Fitzsimons<br />

1965). As largely sessile animals, they rely on<br />

currents to provide them with their edible treats.<br />

Their soft bodies remain out of sight throughout<br />

their lives, slowly adding length to their tubes as<br />

they mature. Their most notable behavior is the<br />

near instantaneous retraction of their radioles back<br />

into the tube when disturbed. The worms can<br />

detect changes in water movement, light intensity<br />

white (Humann & DeLoach 2002), and of course,<br />

tactile stimulation. As such, predators are often<br />

sucker-feeding fish . They have evolved rows of<br />

upper-facing hooks along each segment of their<br />

body, which allow anchoring in the event a fish<br />

attempts to suck them out (Woodin 1987).<br />

Conservation<br />

Although feather dusters themselves are not<br />

seriously threatened, many populations are subject<br />

to ongoing anthropogenic impacts on the coral<br />

reef and mangrove ecosystems, including, but not<br />

limited to, pollution, increased ‘fin’ traffic as<br />

Belizean tourism expands (reef), ocean<br />

acidification (reef), development (mangrove), and<br />

invasive species (Turnball & Harborne 2000).


INVESTIGATION<br />

Introduction<br />

Retraction of radioles is its primary form of<br />

defense, yet they are necessary for both respiration<br />

and ingestion. Given that they are sessile, I would<br />

assume that the animal would achieve the most<br />

benefit from having its radioles extended as long<br />

and often as possible. Therefore, a predator<br />

physically touching the radioles should cause a<br />

full retraction, with a longer latency, versus a<br />

distant water disturbance causing a shorter<br />

retraction and latency. Two tests were devised<br />

for this hypothesis: is there a correlation between<br />

latency and disturbance type; and, do worms<br />

respond differently to tactile and distant<br />

disturbance?<br />

Methods<br />

S. magnifica were identified in two locations<br />

along western shore of Spanish Lookout Caye:<br />

randomly distributed along a concrete seawall and<br />

clustered against square concrete pillars once used<br />

for a pier. Fourteen individuals were found<br />

attached to the seawall, and sixteen were found<br />

attached to different pillars, often in groups of<br />

two. Traditional snorkel gear was used for data<br />

collection, including a floatation device so as to<br />

keep unwanted disturbance at a minimum.<br />

Individuals are disturbed physically by a light<br />

index finger touch. For water movement<br />

disturbance, a ruler was used to gauge a distance<br />

of about 5-6cm from the radioles, and then I<br />

waved my hand laterally until the worm retracted.<br />

Some individuals would not respond to the wave,<br />

in which case I would flick my four fingers from<br />

behind my thumb towards the animal. A static<br />

distance of 5cm was originally intended, though<br />

the sensitive nature of the animal prevented this.<br />

A full or partial retraction was noted. After<br />

retraction, a stopwatch is used to measure the<br />

latency until reemergence. Reemergence from full<br />

retraction is defined as radiole tips protruding<br />

from the rim of its tube, or visible anterior<br />

movement if it is partially retracted.<br />

Results<br />

Using a correlated one-tailed t-test for latency of<br />

each disturbance type, a significant difference was<br />

observed between tactile and distant disturbances in<br />

S. magnifica (t = +2.58, df=29, P < 0.01, n=30),<br />

supporting this portion of the hypothesis. However,<br />

using a 2x2 categorical analysis chi-square test, no<br />

significant difference was found between full and<br />

partial retraction ( 2 = 3.27, df=3, P = 0.3518).<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Discussion<br />

The data collected only partially supports my<br />

hypothesis: tactile disturbance definitely results in<br />

a longer latency than distant water disturbance (P<br />

= 0.007), although full or partial retraction among<br />

individuals appears to be random. An alternate<br />

hypothesis could be that perhaps radioles grow<br />

more rapidly relative to tube construction, and<br />

may protrude even when the animal has retracted<br />

as far as it can. Another possibility is that<br />

individuals can detect the strength of the<br />

disturbance, either tactile or distant, and react<br />

accordingly. Further investigation using measured<br />

static force could be conducted to test this<br />

hypothesis.<br />

References<br />

Humann, P. and N. Deloach. (2002). Reef Creature<br />

Identification, Sea Stars, p. 140 & 148. New<br />

World Publications.<br />

Turnball, C., & Harborne, A. (2000). Summary of<br />

Coral Cay Conservation's Atlantic and Gulf<br />

Rapid Reef Assessment Data From Turneff<br />

Atoll, Belize. London: Coral Cay<br />

Conservation LTD.<br />

Fitzimons, G. (1965). Feeding and Tube-Building in<br />

Sabellastarte magnifica (Shaw). Bulletin of<br />

Marine Science, 15 (3) : 642-670


Marine Hermit Crab<br />

Clibanarius vittatus<br />

Michelle Cassidy<br />

Ecology and Behavior of Manatees and Dolphins in<br />

Belize, Class of 2011<br />

(contact: caryns@sirenian.org)<br />

Taxonomy<br />

Kingdom Animalia<br />

Phylum Arthopoda<br />

Class Crustacea<br />

Order Decapoda<br />

Family p<br />

Genus Clibinarius<br />

Species Clibinarius vittatus<br />

Ecology<br />

Common to most marine, estuarine, and<br />

shoreline environments, hermit crabs can<br />

be easily found worldwide. On Spanish<br />

Lookout Caye in Belize, Central America,<br />

both terrestrial and aquatic varieties of this<br />

crab can be easily spotted; the former found<br />

wandering the sandy shore, and the latter in<br />

the muddy shallows of the mangrove<br />

swamp. The mangrove species, C. vittatus,<br />

feeds on plant and animal detritus that sinks<br />

to the bottom.<br />

Physical Description<br />

Hermit crabs utilize shells (generally from<br />

gastropods) for protection. This results from<br />

their development of a non-calcified<br />

abdomen that would make them vulnerable<br />

to predation (Schejter & Mantelatto 2009).<br />

Crabs have large foreclaws and smaller<br />

hindclaws. Shell sizes and coloration vary<br />

significantly between individuals.<br />

Conservation<br />

Though the hermit crab itself is not<br />

endangered, human expansion threatens its<br />

habitat. These crabs play a key role in the<br />

Mangrove ecosystem via scavenging and<br />

consuming detritus. In order to protect these<br />

animals and the web of life they exist within,<br />

it is important to conserve Mangrove<br />

swamps, such as the ones on Spanish<br />

Lookout Caye. Plenty of Mangrove Cayes in<br />

Belize have been dredged and filled to<br />

provide land for resorts, the process of<br />

which destroys valuable Mangrove<br />

resources and hermit crab habitats.<br />

Promoting awareness of the numerous<br />

species that depend on Mangroves for<br />

survival is key to conserving these animals.<br />

Behavior<br />

Crabs retreat quickly into their shells at any<br />

sign of potential danger. They will warily<br />

emerge from their shells if not touched or<br />

approached for approximately one minute.<br />

Competition for particularly nutirient-rich<br />

patches can become aggressive, with larger<br />

crabs often strong-arming smaller<br />

individuals into leaving the area, so they can<br />

enjoy the food source (Ramsay et al. 1997).<br />

Hermit crabs reproduce seasonally, with<br />

intensive mating activity occurring in the<br />

warmer months. Studies show that the<br />

recruitment area for young crabs differs<br />

from the adult habitat, and individuals are<br />

already an intermediate size once they<br />

appear. (Sant´Anna et al. 2008, Turra &<br />

Leite 2000)<br />

Crabs can detect chemical signals of a dead<br />

or dying gastropod or conspecific, alerting<br />

them to the presence of a potential home<br />

(Rittschof 1980, Rittschof et al. 1992). They<br />

can also use chemical stimuli to identify<br />

predators (Rosen et al. 2009).


Introduction<br />

Shells serve as a limiting factor for hermit<br />

crabs in most environments (Tricarico &<br />

Gherardi 2006). As a result, crabs must<br />

develop strategic behaviors to procure<br />

viable shells as they grow. Synchronous<br />

vacancy chains (SVC) are adaptive<br />

behaviors that address this limitation by<br />

providing SVCs have been observed in<br />

terrestrial hermit crabs in the Drowned<br />

Cayes of Belize (Rotjan et al. 2010), but it is<br />

unknown if the marine species C. vittatus<br />

also implements SVCs. This study aims to<br />

answer the question of whether marine<br />

hermit crabs also utilize SVCs for shell<br />

acquisition.<br />

Closely related species of terrestrial Hermit Crabs<br />

(Coenobita clypeatus) in a SVC on SLC.<br />

Materials and Methods<br />

I gathered medium and large empty<br />

gastropod shells from the seagrass beds<br />

near the mangrove pond, then placed them<br />

within four 30 cm x 30 cm quadrats placed<br />

within the hermit crab habitat. I observed for<br />

two hour and a half sessions, making note<br />

of behavior in each quadrat every 10<br />

minutes. The number of hermit crabs<br />

present and the presence or absence of<br />

characteristic SVC behaviors were noted<br />

(piggybacking, waiting, tug of war). A SVC<br />

was present if the largest crab moves into<br />

an empty shell, while smaller crabs lined up<br />

behind move successively into the shell of<br />

the crab before it.<br />

Results<br />

Data collected couldnt be statistically<br />

analyzed, but some patterns emerged from<br />

observation. Crabs investigated empty<br />

shells, and one individual waited for<br />

approximately 30 minutes by a shell that<br />

was too large for it. On several occasions,<br />

crabs would “piggyback”, grasping to the<br />

shell of a larger individual while it moved<br />

around the area. The was one incident of<br />

“tug of war” in which two crabs pulled an<br />

empty shell back and forth between them<br />

for approximately ten minutes. Eventually<br />

the smaller crab walked away, and the<br />

larger crab switched into the empty shell.<br />

No other crabs came to take the shell left<br />

behind by that switch.<br />

Discussion<br />

While no SVCs were observed for C.<br />

vittatus, there was evidence of behaviors<br />

such as piggybacking and waiting that<br />

suggest that this species of hermit crab also<br />

utilizes this method of shell acquisition. It is<br />

possible that for this population, shells are<br />

not as vital as food—many individuals in the<br />

study area ignored the shells completely<br />

and spent their time feeding. The one shell<br />

switch I witnessed suggests that both<br />

synchronous and asynchronous vacany<br />

chains are useful methods of gaining new<br />

protective shells. It would require further<br />

investigation to evidence which method<br />

occurs more frequently for the marine hermit<br />

crabs of Spanish Lookout Caye.<br />

Literature Cited<br />

Rotjan D, Chabot J, and Lewis S. Social<br />

context of shell acquisition in Coenobita<br />

clypeatus hermit crabs. Behavioral Ecology.<br />

(2010) 21: 639-45.<br />

Tricarico E and Gherardi F. Shell acquisition<br />

by hermit crabs: which tactic is more<br />

efficient? Behavioral Ecology Sociobiology.<br />

(2006) 60: 492-500.


Variegated Sea Urchin<br />

Lytechinus variegatus<br />

Kalie Bishop<br />

Ecology and Behavior of Manatees and<br />

Dolphins in<br />

Belize, Class of 2011 (contact:<br />

cselfsullivan@sirenian.org)<br />

Taxonomy<br />

Kingdom: Anamalia<br />

Phylum: Echinodermata<br />

Class: Echinoidea<br />

Order: Temnopleuroida<br />

Family: Toxopneustida<br />

Genus: Lytechinus<br />

Species: Lytechinus variegates<br />

Ecology<br />

L. variegatus is commonly found in shallow,<br />

calm waters, such as seagrass beds. They<br />

have adapted to survive in many marine<br />

environments but primarily subsist within<br />

seagrass and kelp beds. L. variegatus<br />

inhabits the waters from North Carolina and<br />

Bermuda southward to the Caribbean and<br />

Brazil (Hendler et al. 1995). Although they<br />

cannot live in freshwater, they inhabit all<br />

depths and climates of the oceans.<br />

Physical Description<br />

L. variegatus vary in color, but the most<br />

frequently seen are green, purple and light<br />

pink. L. variegatus have radially symmetrical<br />

bodies divided into five equal parts. The<br />

skeletal structure of the sea urchin is a rigid<br />

test or theca that is made up of plates<br />

encircling the mouth in the center of the oral<br />

side, encompassing the urchins inner<br />

organs (Nichols 1962). Spines, which can<br />

regenerate, cover the entire surface of the<br />

sea urchin test and play a role in protection<br />

as well as movement. Among the spines are<br />

five paired rows of tiny tube feet with<br />

suckers that help with locomotion, capturing<br />

food, and holding onto the seafloor.<br />

White variation of L. variegatus<br />

Behavior<br />

L. variegatus feed on decaying matter, kelp,<br />

sea grass, algae and occasionally even sea<br />

urchin skeletons. Sea urchins reproduce by<br />

releasing the unfertilized eggs and sperm<br />

into the water column, where the eggs are<br />

fertilized and develop into free-swimming<br />

larvae known as plutei (Pechenik 2000;<br />

Harvey 1956). The larval form<br />

of L. variegatus, as with most echinoderms,<br />

is very different from the adult form and<br />

must undergo a complex metamorphosis to<br />

reach adult form.<br />

Conservation<br />

Being primary grazers of kelp, and algae, L.<br />

variegatus play an important role within the<br />

marine communities that they inhabit. The<br />

population of L. variegatus in an area is vital<br />

to the health of the coral reef, as they<br />

eliminate the seaweed and algae that can<br />

consume coral if a balance is not<br />

maintained within the environment. The<br />

opposite can also occur with an increase in<br />

L. variegatus causing erosion on coral reef.<br />

It is crucial to have balance within the<br />

environment to maintain healthy<br />

populations, such as the population on<br />

Spanish lookout Caye.


Introduction<br />

Lunar reproductive rhythms have been part<br />

of fishermans folklore since ancient times.<br />

As described in much detail by H. Monroe<br />

Fox in 1923, “Sizes of certain marine<br />

invertebrates, chiefly mollusks and<br />

echinoderms varies with the phases of the<br />

moon.” On a field biology course studying<br />

the Drowned Cayes, I noticed that L.<br />

variegatus was in abundance along the<br />

seawall on the North facing aspect of<br />

Spanish lookout Caye off the coast of Belize<br />

City, Belize. The objective of my research<br />

is to determine if L. variegatus mass<br />

increases in relation to the lunar cycle.<br />

Methods<br />

I selected the North facing concrete seawall<br />

on the West side of Spanish Lookout Caye.<br />

On 08/07/2011, 80/09,2011, and<br />

08/10/2011, I removed 14 individuals from<br />

the area by placing them in a bucket filled<br />

with seawater. All debris was removed from<br />

each individual by hand, then removed from<br />

the bucket for 15 seconds then placed on a<br />

scale to record each individuals body mass.<br />

Individual photos were taken and were<br />

reviewed later for photo identification in<br />

chronological order. Due to technical<br />

difficulties, I was unable to analyze<br />

individuals with photo I.D. from 08/10/2011.<br />

Results<br />

Although there was a trend there was no<br />

significant increase in body mass between<br />

day one and day two. (t=1.67), (df=4),<br />

(p=0.08). Despite failure to meet statistical<br />

criteria, each individual increased slightly<br />

from day one to day two. This suggests that<br />

it is plausible that increase in body mass of<br />

L. variegatus correlates to the lunar cycle.<br />

Green variation of L. variegatus<br />

Discussion<br />

Further investigation should been<br />

conducted using a larger sample size.<br />

Photo identification methods should be<br />

performed with the same high pixel lens<br />

throughout the study to ensure photo<br />

identification methods are correct, as this<br />

skewed my data and conclusions. Further<br />

research should be conducted throughout<br />

the entire duration of the lunar cycle, from<br />

new moon until full moon.<br />

Literature Cited<br />

Hendler, G., Kier, P.M., Miller, J.E.,<br />

Pawson, D.L. 1995. Sea Stars, Sea<br />

Urchins, and Allies: Echinoderms of<br />

Florida and the Caribbean. Washington:<br />

Smithsonian Institution Press; 390p<br />

Nichols, D. 1962. Echinoderms. London:<br />

Hutchinson & CO LTD; 192p.<br />

Pechenik, J.A. 2000. Biology of the<br />

Invertebrates, Fourth Edition. U.S.A.:<br />

McGraw-Hill Companies, Inc.; 578p.


Our Data Sheets<br />

We have designed our data sheets for long-term project use and built in redundancy to<br />

ensure we don’t miss data. They have been field tested over the past decade with high<br />

success.<br />

We organize the data sheets in field bins. Each bin should have sharpened pencils, a<br />

white eraser, and datasheets as described below. Please take the time to review these<br />

datasheets before fielding!<br />

We will use 5 bins total. Each volunteer/student will be responsible for one bin per day.<br />

After they master their data sheet, they will rotate jobs and teach each other, with<br />

oversight from the PI, Visiting Scientists, TAs, and Interns. The PI will personally teach<br />

and monitor for the first few days. If all goes as expected, interns and TAs will have<br />

mastered the forms with a few days in the field.<br />

One bin will have the Trip Summary (2-page data sheet). We should have 2 page-one<br />

of the Trip Summary per day (~20 per team) and 10 page-two per day (~100 per team).<br />

The team member who is responsible for this form is writing pretty much all day.<br />

The 2nd bin will have the Record of Effort (1-page data sheet) in it. We should have 4<br />

(front and back for a total 8) Record of Effort Sheets per day (~40/team). The team<br />

member who is responsible for this form is writing pretty much all day.<br />

The 3rd bin will have the Manatee Sighting and Scan (4-page data sheet). We should<br />

have 10 sets/day (~100/team). The team member who is responsible for this form is only<br />

writing during a scan or an opportunistic manatee sighting.<br />

The 4th bin will have the Habitat Sampling (4-page data sheet). We should have 10<br />

sets/day (~100/team). The team member who is responsible for this form is only writing<br />

during the habitat sample, which follows every scan.<br />

The 5th bin will have the Dolphin Survey (2-page data sheet). We should have 5<br />

sets/day (~50/team). The team member who is responsible for this form is documenting<br />

when we depart and return, but only writing down data when we encounter dolphins.<br />

Copies of our study site (map), marine life and bird field cards should also be kept in the<br />

bins.


TRIP SUMMARY SHEET (page 1) Day of the Week:<br />

Trip ID: Date: Julian Day:<br />

(yy-julianday-onedigittrip#) (dd-mon-yy) (001-365)<br />

EW Team #: # EW Vols: Total Obs:<br />

(yy-one digit team#)<br />

EW Team Members (first and last name):<br />

Researcher(s): Intern(s): Field Asst.:<br />

(first name) (firstname and initials) (firstname and initials)<br />

Effort Data Taken by:<br />

Sighting/Scan Data Taken by:<br />

Trip Summary Data Taken by:<br />

Behavioral Data Taken by:<br />

Seagrass Data Taken by:<br />

Locations (scan locations): ________________________________________________________<br />

_________________________________________________________________________<br />

Total # of Scans: Total # of Sightings (opp/scan): Total # Manatees:_______<br />

GENERAL WEATHER CONDITIONS<br />

DEPART DOCK: WEATHER CHECK I : WEATHER CHECK II:<br />

RETURN DOCK: Time: Time:<br />

*TIDES IN ORDER: Wind Speed: mph Wind Speed: mph<br />

High: Wind Direction: Wind Direction:<br />

Low: Sea State: Sea State:<br />

High: Swell Height: ft Swell Height: ft<br />

Low: Cloud Cover: Cloud Cover:<br />

High: Water Temp: °C Water Temp: °C<br />

Low: Air Temp: °C Air Temp: °C<br />

Rainfall Overnight: Salinity: PPT Salinity: PPT<br />

Rainfall Today: Barometer: HPa Barometer: HPa<br />

Cruise Ships (#):<br />

Comments:<br />

* If a high or low falls pre/post today’s date, put in parentheses in order of occurrence.<br />

Please Complete Trip Log on reverse side and additional pages!<br />

CSS/KSL ____________ Trip Summary Sheet Page_______Of_______<br />

(c) 2001-2004 Manatees in Belize - Earthwatch Project PIs: Katherine S. Lacommare & Caryn Self Sullivan -Updated 4/7/11


TRIP SUMMARY - LOG SHEET Date Page _____ of _____<br />

Record All Events during the Day<br />

TIME WPoint EVENT LOCATION LATITUDE LONGITUDE COMMENTS<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

<strong>17</strong>. 88.<br />

2001-2004 Manatees in Belize <strong>–</strong> Earthwatch Project PIs: Katherine S. LaCommare and Caryn Self Sullivan 4/7/2011


RECORD OF EFFORT DATA SHEET<br />

Event (include num.) Location or Route from/to: ____________________________<br />

Enter one of the following: Travel (trav)<strong>–</strong> Survey (surv)<strong>–</strong> Pt. Scan (pscn) <strong>–</strong> Sighting (sght)<strong>–</strong> Sampling (samp)- Other Work (othw) <strong>–</strong> Lunch (lunc)<strong>–</strong> Other Free (othf)<br />

Start time _________Stop time _______Average Speed (km/hr) Trip (km)______________<br />

Start WP (Garmin) _________ Stop WP (Garmin) ________EPE = _ # Manatees__________<br />

Comments:<br />

Event (include num.) Location or Route from/to: ____________________________<br />

Enter one of the following: Travel (trav)<strong>–</strong> Survey (surv)<strong>–</strong> Pt. Scan (pscn) <strong>–</strong> Sighting (sght)<strong>–</strong> Sampling (samp)- Other Work (othw) <strong>–</strong> Lunch (lunc)<strong>–</strong> Other Free (othf)<br />

Start time _________Stop time _______Average Speed (km/hr) Trip (km)_____________<br />

Start WP (Garmin) _________ Stop WP (Garmin) ________EPE = __ # Manatees_________<br />

Comments:<br />

Event (include num.) Location or Route from/to: ____________________________<br />

Enter one of the following: Travel (trav)<strong>–</strong> Survey (surv)<strong>–</strong> Pt. Scan (pscn) <strong>–</strong> Sighting (sght)<strong>–</strong> Sampling (samp)- Other Work (othw) <strong>–</strong> Lunch (lunc)<strong>–</strong> Other Free (othf)<br />

Start time _________Stop time _______Average Speed (km/hr) Trip (km)_____________<br />

Start WP (Garmin) _________ Stop WP (Garmin) ________EPE =__ # Manatees_________<br />

Comments:<br />

Event (include num.) Location or Route from/to: ____________________________<br />

Enter one of the following: Travel (trav)<strong>–</strong> Survey (surv)<strong>–</strong> Pt. Scan (pscn) <strong>–</strong> Sighting (sght)<strong>–</strong> Sampling (samp)- Other Work (othw) <strong>–</strong> Lunch (lunc)<strong>–</strong> Other Free (othf)<br />

Start time _________Stop time _______Average Speed (km/hr) Trip (km)_____________<br />

Start WP (Garmin) _________ Stop WP (Garmin) ________EPE = __ # Manatees_________<br />

Comments:<br />

DATE: ______________________ Effort Data Taken By? _____________________<br />

(dd-mon-yy)<br />

Trip ID: _____________________ RECORD OF EFFORT Page _______of________<br />

(yy-julianday-one digit trip#)<br />

(c) 2001-2007 Manatees in Belize - Earthwatch Project PIs: Caryn Self Sullivan & Katie LaCommare -Updated 4/7/11


MANATEE SIGHTINGS AND SCANS Page 1 of ____ ID # _______________________________<br />

(yy-julianday-sight#scan#)<br />

Sighting Number Sight Start (24 hr) Sight Stop<br />

Point Scan Number Scan Start (24 hr) Scan Stop<br />

Sighting Type (circle one): Opportunistic or Scan<br />

Wayppoint (Garmin): EPE:<br />

Location: Loc. Code (see list for proper code):<br />

MUST Enter for both sightings and scans ONLY Enter code if this is a scan (whether there is a sighting or not).<br />

# Of Other Boats in area during scan: Distance (closest point) (1) (2) ___(3)<br />

Comments (size, type, speed) (1) (2) (3)<br />

Manatee Sighted By__________________ Confirmed By________________________<br />

Enter one of the following: Researcher, Field Assistant, Intern, Volunteer, Team, Not confirmed<br />

*Total # Of Manatees No. Calves Calf Size NEW, YOY, OTHER<br />

1 ST 20 minutes of SCAN: Minimum: Maximum: Best Estimate: *TOTAL # DURING SIGHTING<br />

Total 30 minutes of SCAN: Minimum: Maximum: Best Estimate:<br />

__________________________________________________________________________________________<br />

Initial Distance to 1st Manatee at 1 st Sighting (m): Time: Initial Movement of 1st Manatee:<br />

Aw from boat Towards boat Milling No change Undetermined<br />

(CIRCLE Distance Detection Method: DIRECT INDIRECT ESTIMATE) (CIRCLE initial movement in relationship to observation boat)<br />

Habitat Type: (Circle habitat where the manatee is, NOT the scan pt. ) UNDETERMINED Resting hole-yes or no or unk<br />

InBogue/OutBogue/Reef | Chann/ChannEdge/DeadEndBogue/Lagoon/Grassflat/UNK | Mud/Grass/Sand/Coral/UNK | Turtle/Shoal/Manatee/None/UNK<br />

DISTURBED? YES OR NO OR UNK Predominate Behavioral State (circle one):<br />

Feeding Resting Socializing Traveling Milling Undetermined Other (describe):<br />

__________________________________________________________________________________________<br />

Initial Distance to 2 nd Manatee at 1 st Sighting (m): Time: Initial Movement of 2 nd Manatee:<br />

Aw from boat Towards boat Milling No change Undetermined<br />

(CIRCLE Distance Detection Method: DIRECT INDIRECT ESTIMATE) (CIRCLE initial movement in relationship to observation boat)<br />

Habitat Type: (Circle habitat where the manatee is, NOT the scan pt. ) UNDETERMINED Resting hole-yes or no or unk<br />

InBogue/OutBogue/Reef | Chann/ChannEdge/DeadEndBogue/Lagoon/Grassflat/UNK | Mud/Grass/Sand/Coral/UNK | Turtle/Shoal/Manatee/None/UNK<br />

DISTURBED? YES OR NO OR UNK Predominate Behavioral State (circle one):<br />

Feeding Resting Socializing Traveling Milling Undetermined Other (describe):<br />

__________________________________________________________________________________________<br />

Initial Distance to Center of Manatee Group at 1 st Sighting (m):____________________________________<br />

U/W Video Attempt : yes no; if yes, was it successful? _____ Seagrass Sample/Habitat Sample/No Sample<br />

SIGHTING/SCAN CONDITIONS<br />

Glare<br />

Yes<br />

No<br />

Glare<br />

Direction<br />

Cloud cover<br />

clear<br />

scattered<br />

partly<br />

mostly<br />

overcast<br />

Precipitation<br />

Dry<br />

Light rain<br />

Heavy rain<br />

Sea State<br />

Swell Height Tide State*<br />

High (+/- 60m)<br />

Low (+/- 60m)<br />

Flood Ebb<br />

* If Scan Start Time falls within 60 min of High or Low Tide, then circle High or Low; if not and Scan Start Time falls b/t High and Low, circle Ebb; if not and Scan<br />

Start Time falls b/t Low and High, circle Flood.<br />

DATE: ______________________ Sighting/Scan Data Taken By? ______________<br />

(dd-mon-yy)<br />

Trip ID: _____________________ SIGHTINGS AND SCANS Data Set _______of________<br />

(yy-julianday-one digit trip#)<br />

(c) 2001-2004 Manatees in Belize - Earthwatch Project PIs: Katherine S. Lacommare & Caryn Self Sullivan -Updated 4/7/11


MANATEE SIGHTINGS AND SCANS Page 2 of ____ ID # _______________________________<br />

(yy-julianday-sight#scan#)<br />

Site Map: On this page you will sketch the characteristics of the scan point in relationship to the<br />

Boat. Map Perspective: illustrator is standing on the driver’s seat and facing north when<br />

looking towards the top of this page. Sketch the above water characteristics and draw a vector<br />

(distance & direction) to each landmark and to each manatee sighted. Label each manatee vector<br />

with time, distance, & direction.<br />

N<br />

W E<br />

S<br />

(c) 2001-2004 Manatees in Belize - Earthwatch Project PIs: Katherine S. Lacommare & Caryn Self Sullivan -Updated 4/7/11


MANATEE SIGHTINGS AND SCANS Page 3 of ____ ID # _______________________________<br />

(yy-julianday-sight#scan#)<br />

Comments:_________________________________________________________________________________<br />

___________________________________________________________________________________________<br />

CatalogID # _______________________ Marked / Un-Marked / Undetermined Sex: M / F / U Prop Scars? Yes / No<br />

RESIGHT? _______________________ Tape Start _______Stop______ Animal _____ of _____ captured on tape<br />

Enter Feature Codes* 2 (see below)<br />

for each unique marking seen on animal<br />

__ __ __ __ __ __ __ __ __ __<br />

__ __ __ __ __ __ __ __ __ __<br />

__ __ __ __ __ __ __ __ __ __<br />

__ __ __ __ __ __ __ __ __ __<br />

__ __ __ __ __ __ __ __ __ __<br />

__ __ __ __ __ __ __ __ __ __<br />

__ __ __ __ __ __ __ __ __ __<br />

__ __ __ __ __ __ __ __ __ __<br />

__ __ __ __ __ __ __ __ __ __<br />

Barnacles: Yes / No / Undetermined<br />

Number? ______ Size? ____________<br />

Remoras: Yes / No / Undetermined<br />

Number? ________________________<br />

Algae: Yes / No / Undetermined<br />

Color/Coverage? __________________<br />

Medial Notch: Yes / No / Undetermined<br />

Comments? ______________________<br />

Behavioral State: Rest / Feed / Travel / Social / Mill<br />

Play / Other / Undetermined<br />

Initial Reaction to diver: Approach / Retreat / No Change / Touch / Other<br />

Secondary Reaction to diver: Approach / Retreat / No Change / Touch / Other<br />

* 2 Identifying Feature Codes to use in blanks above, use one line for each unique marking:<br />

Type Region Position Number Size Color Shape<br />

S <strong>–</strong> scar D <strong>–</strong> dorsal F <strong>–</strong> flipper 1 <strong>–</strong> single L <strong>–</strong> large G <strong>–</strong> gray B <strong>–</strong> blotch<br />

M <strong>–</strong> mutilation L <strong>–</strong> left H <strong>–</strong> head 2 <strong>–</strong> 2 or 3 M <strong>–</strong> medium W <strong>–</strong> white L <strong>–</strong> line(s)<br />

D <strong>–</strong> deformity R <strong>–</strong> right A <strong>–</strong> ant. trunk 4 <strong>–</strong> 4 or more S <strong>–</strong> small<br />

F <strong>–</strong> freeze brand V <strong>–</strong> ventral B <strong>–</strong> mid. trunk<br />

N <strong>–</strong> medial notch C <strong>–</strong> post. trunk<br />

K <strong>–</strong> trunk plain D <strong>–</strong> peduncle<br />

L <strong>–</strong> tail plain X <strong>–</strong> ant. tail<br />

P <strong>–</strong> skin pigmentation Y <strong>–</strong> post. tail<br />

(c) 2001-2004 Manatees in Belize - Earthwatch Project PIs: Katherine S. Lacommare & Caryn Self Sullivan -Updated 4/7/11


MANATEE SIGHTINGS AND SCANS Page 4 of ____ ID # _______________________________<br />

(yy-julianday-sight#scan#)<br />

__________________________________________________________________________________________<br />

Initial Distance to 3 rd Manatee at 1 st Sighting ((m): Time: Initial Movement of 3 rd Manatee:<br />

Aw from boat Towards boat Milling No change Undetermined<br />

(CIRCLE Distance Detection Method: DIRECT INDIRECT ESTIMATE) (CIRCLE initial movement in relationship to observation boat)<br />

Habitat Type: (Circle habitat where the manatee is, NOT the scan pt. ) UNDETERMINED Resting hole-yes or no or unk<br />

InBogue/OutBogue/Reef | Chann/ChannEdge/DeadEndBogue/Lagoon/Grassflat/UNK | Mud/Grass/Sand/Coral/UNK | Turtle/Shoal/Manatee/None/UNK<br />

DISTURBED? YES OR NO OR UNK Predominate Behavioral State (circle one):<br />

Feeding Resting Socializing Traveling Milling Undetermined Other (describe):<br />

__________________________________________________________________________________________<br />

Initial Distance to 4th Manatee at 1 st Sighting ((m): Time: Initial Movement of 4th Manatee:<br />

Aw from boat Towards boat Milling No change Undetermined<br />

(CIRCLE Distance Detection Method: DIRECT INDIRECT ESTIMATE) (CIRCLE initial movement in relationship to observation boat)<br />

Habitat Type: (Circle habitat where the manatee is, NOT the scan pt. ) UNDETERMINED Resting hole-yes or no or unk<br />

InBogue/OutBogue/Reef | Chann/ChannEdge/DeadEndBogue/Lagoon/Grassflat/UNK | Mud/Grass/Sand/Coral/UNK | Turtle/Shoal/Manatee/None/UNK<br />

DISTURBED? YES OR NO OR UNK Predominate Behavioral State (circle one):<br />

Feeding Resting Socializing Traveling Milling Undetermined Other (describe):<br />

__________________________________________________________________________________________<br />

Sighting Comments (con’t):<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

____________________________________________________________________<br />

(c) 2001-2004 Manatees in Belize - Earthwatch Project PIs: Katherine S. Lacommare & Caryn Self Sullivan -Updated 4/7/11


Date: Page 1 of ____ Sight/Scan Record ID:<br />

(dd-mon-yy) (yy-julianday-sighting#scan#)<br />

HABITAT SAMPLING<br />

Sample ID: Data Recorded By:<br />

(yy-julianday-location code)<br />

LOCATION DATA<br />

LOC. CODE: WAYPOINT: LOCATION NAME:<br />

PHYSICAL DATA<br />

SECCHI READINGS (vertical secchi <strong>–</strong> take from center of boat; horizontal secchi <strong>–</strong> take from bow (0.5 meter)<br />

below the surface; secchi disk should face the sun.)<br />

Name (volunteer who took the data) Horiz. Vertical Depth Comments<br />

Air Temp<br />

°C<br />

H20 Temp<br />

Surface °C<br />

H20 Temp<br />

Bottom °C<br />

Salinity<br />

Surface ‰<br />

Salinity<br />

Bottom ‰<br />

Sea State<br />

Beaufort<br />

Swell<br />

Height (ft)<br />

Habitat Type: Circle the habitat characteristics of the scan point (where the boat is, NOT where the manatee<br />

was observed). This may be different from where the manatee was sighted or where the plot was set.<br />

InBogue/OutBogue/Reef | Chann/ChannEdge/DeadEndBogue/Lagoon/Cove/Grassflat<br />

Mud/Sandy Mud/Muddy Sand/ Sand/Coral | Turtle/Shoal/Manatee/None<br />

Comments and other description:___________________________________________________________________<br />

______________________________________________________________________________________________<br />

______________________________________________________________________________________________<br />

PLOT DATA<br />

VECTOR (Distance in yards & Direction N, NE, E...) TO SAMPLE PLOT:<br />

Habitat Type: Circle the habitat characteristics of the plot NOT where the boat is but where the plot is. This<br />

may be different from where the boat is.<br />

InBogue/OutBogue/Reef | Chann/ChannEdge/DeadEndBogue/Lagoon/Cove/Grassflat | Mud/Sand/Coral | Turtle/Shoal/Manatee/None<br />

PLOT DEPTH at poles -- Center: North: South: East: West:<br />

Sediment Type (circle sediment type based on biomass cores): mud -- sand/coral <strong>–</strong> mud/sand -- -- other<br />

2001 Manatees in Belize <strong>–</strong> Earthwatch Project PIs: Katherine S. LaCommare and Caryn Self Sullivan 4/7/2011


Date: Page 2 of ____ Sight/Scan Record ID:<br />

(dd-mon-yy) (yy-julianday-sighting#scan#)<br />

BIOLOGICAL DATA<br />

BIOMASS CORES: (Enter zero if core would be empty; enter yes if core was taken; enter N/A if not taken. Take<br />

cores from the “a” corner of the small (density) quadrats. If you enter zero, a biomass datasheet must be filled out!!! This<br />

does not need to be entered into the database)<br />

N:__________S:__________E:__________W:__________<br />

DENSITY COUNTS<br />

4 samples:<br />

See plot design for sample layout.<br />

At least 2 Replicas: If possible, each replica should be taken by the same person. Please make comments about person<br />

who took data i.e., intern, volunteer etc.)<br />

Replica 1: Name (volunteer who took the data):<br />

A quarter B quarter C quarter D quarter Total<br />

N: T/ S/ M T/ S/ M T/ S/ M T/ S/ M T/ S/ M<br />

S: T/ S/ M T/ S/ M T/ S/ M T/ S/ M T/ S/ M<br />

E: T/ S/ M T/ S/ M T/ S/ M T/ S/ M T/ S/ M<br />

W: T/ S/ M T/ S/ M T/ S/ M T/ S/ M T/ S/ M<br />

Replica 2: Name (volunteer who took the data):<br />

A quarter B quarter C quarter D quarter Total<br />

N: T/ S/ M T/ S/ M T/ S/ M T/ S/ M T/ S/ M<br />

S: T/ S/ M T/ S/ M T/ S/ M T/ S/ M T/ S/ M<br />

E: T/ S/ M T/ S/ M T/ S/ M T/ S/ M T/ S/ M<br />

W: T/ S/ M T/ S/ M T/ S/ M T/ S/ M T/ S/ M<br />

Replica 3: Name (volunteer who took the data):<br />

A quarter B quarter C quarter D quarter Total<br />

N: T/ S/ M T/ S/ M T/ S/ M T/ S/ M T/ S/ M<br />

S: T/ S/ M T/ S/ M T/ S/ M T/ S/ M T/ S/ M<br />

E: T/ S/ M T/ S/ M T/ S/ M T/ S/ M T/ S/ M<br />

W: T/ S/ M T/ S/ M T/ S/ M T/ S/ M T/ S/ M<br />

2001 Manatees in Belize <strong>–</strong> Earthwatch Project PIs: Katherine S. LaCommare and Caryn Self Sullivan 4/7/2011


Date: Page 3 of ____ Sight/Scan Record ID:<br />

(dd-mon-yy) (yy-julianday-sighting#scan#)<br />

PERCENT COVER (1 x 1 m QUADRAT)<br />

4 samples N,S,E,W: See plot design for sample layout. Each quadrat will be rolled 5 times.<br />

Replica 1: Name (volunteer who took the data):<br />

Sample N (grand total is entered into database) (rows 1-10)<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Total<br />

Grand Total<br />

Replica 1: Name (volunteer who took the data):<br />

Sample S (grand total is entered into database) (rows 1-10)<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Total<br />

Grand Total<br />

Replica 1: Name (volunteer who took the data):<br />

Sample E (grand total is entered into database) (rows 1-10)<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Total<br />

Grand Total<br />

Replica 1: Name (volunteer who took the data):<br />

Sample W (grand total is entered into database) (rows 1-10)<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Total<br />

Grand Total<br />

2001 Manatees in Belize <strong>–</strong> Earthwatch Project PIs: Katherine S. LaCommare and Caryn Self Sullivan 4/7/2011


Date: Page 4 of ____ Sight/Scan Record ID:<br />

(dd-mon-yy) (yy-julianday-sighting#scan#)<br />

Replica 2: Name (volunteer who took the data):<br />

Sample N (grand total is entered into database) (rows 1-10)<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Total<br />

Grand Total<br />

Replica 2: Name (volunteer who took the data):<br />

Sample S (grand total is entered into database) (rows 1-10)<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Total<br />

Grand Total<br />

Replica 2: Name (volunteer who took the data):<br />

Sample E (grand total is entered into database) (rows 1-10)<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Total<br />

Grand Total<br />

Replica 2: Name (volunteer who took the data):<br />

Sample W (grand total is entered into database) (rows 1-10)<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Total<br />

Grand Total<br />

2001 Manatees in Belize <strong>–</strong> Earthwatch Project PIs: Katherine S. LaCommare and Caryn Self Sullivan 4/7/2011


DOLPHIN ENCOUNTER: Boat-Based Data Sheet Location: Drowned Cayes, Belize, Central America<br />

Survey # Encounter # Date: 20____ Boat: Captain: Photographer: # of images:<br />

(day month)<br />

Survey Start Time: End Time: Effort: Hours with Dolphins: Sunrise: Sunset: Moonrise: Moonset:<br />

Tide 1stHigh: 1stLow: High: Low: High: Total Survey Effort to Date: Total Hours with Dolphins to Date: Total Images to Date:<br />

Start Initial GPS Reading Sea Water Water Sighting Description Identified IDs:<br />

Time Tide State Swell Depth Temp Area Habitat Other Objects<br />

N W (m) (m) (C) General Comments:<br />

End Final GPS Reading Sea Water Water<br />

Sighting Description<br />

Time Tide State Swell Depth Temp Area Habitat Other Objects<br />

N W (m) (m) (C)<br />

Time Group Estimate ID Ratio<br />

Group Membership Marked:<br />

Calves Juveniles Adults Unmarked:<br />

< 1 month 1 - 3 mos 3 - 6 mos 6 - 10 mos Total < 1 yr 1 - 2 yr Total:<br />

Best Estimate ID-Ratio:<br />

SIGHTING LOG<br />

Way GPS Reading Area / Habitat Water Group Behavior Response to<br />

Time Pnt. N W Boats / Birds / Other depth Size State Research Vessel Notes and Comments<br />

Contact Details: Dr. Caryn Self-Sullivan, Ph.D., <strong>Sirenian</strong> <strong>International</strong>, 200 Stonewall Drive, Fredericksburg, VA 22401 Email: caryns@sirenian.org; Phone: 540.287.8207


Sighting Log Continued from Page 1<br />

Way GPS Reading Area / Habitat Water Group Behavior Response to<br />

Time Pnt. N W Boats / Birds / Other depth Size State Research Vessel Notes and Comments<br />

Contact Details: Dr. Caryn Self-Sullivan, Ph.D., <strong>Sirenian</strong> <strong>International</strong>, 200 Stonewall Drive, Fredericksburg, VA 22401 Email: caryns@sirenian.org; Phone: 540.287.8207


Caryn Self-Sullivan, Ph.D.<br />

200 Stonewall Drive, Fredericksburg, VA 22401-2110<br />

cselfsullivan@gmail.com | 540.287.8207<br />

Education<br />

Ph. D. 2008 Texas A&M University, College Station, TX, Department of Wildlife and Fisheries Sciences<br />

B. S. 1997 Coastal Carolina University, Conway, Department of Marine Science, minors in Biology and<br />

Mathematics<br />

2000 Pennsylvania State University, Graduate Program in Acoustics, non-degree seeking student (June)<br />

1996 Deakin University, Geelong, Victoria, Australia, exchange student (July-December)<br />

Professional Positions<br />

2011-Present Distance-Education Faculty (Adjunct), Fischler School of Education and the Oceanographic Center,<br />

Nova Southeastern University<br />

2010-2011 Assistant Professor (visiting, full-time faculty) Department of Fisheries & Wildlife, Virginia Tech<br />

2008-2010 Assistant Professor (temporary, full-time faculty) Department of Biology, Georgia Southern University<br />

2008 Substitute Teacher (Secondary Level) for Spotsylvania, Stafford, Fredericksburg, VA<br />

2010-Present Technical Expert to Friends of Swallow Caye, Swallow Caye Wildlife Sanctuary, Belize<br />

2010-Present Technical Expert to the Sarteneja Allicance for Conservation & Development, Corozol Bay Wildlife<br />

Sanctuary, Belize<br />

2007-Present Marine Science Advisor, Hugh Parkey Foundation for Marine Awareness & Education, Belize<br />

2005-Present Principal Advisor, NCRC Community Conservation of West African Manatees in Volta Lake, Ghana<br />

2005-Present Member, IUCN Species Survival Commission Sirenia Specialist Group<br />

2004-Present Member, Belize National Manatee Working Group<br />

2004-Present Member, Belize Marine Mammal Stranding Network<br />

2000-Present President & Co-founder, <strong>Sirenian</strong> <strong>International</strong>, Inc.<br />

1998-Present Principal Investigator, Ecology and Behavior of Antillean Manatees and Bottlenose Dolphins in the<br />

Drowned Cayes Area of Belize<br />

2004-2007 Director of Research & Education, Spanish Bay Conservation & Research Center and the Hugh Parkey<br />

Foundation for Marine Awareness and Education, Belize<br />

Professional Memberships: National Science Teachers Association (NSTA), North American Association for<br />

Environmental Education (NAAEE), Society for Marine Mammalogy (SMM), Society for<br />

Conservation Biology (SCB), Animal Behavior Society (ABS), American Pet Dog Trainers<br />

Association (APDT), Karen Pryor Academy Certified Training Partner (KPA CTP), <strong>Sirenian</strong><br />

<strong>International</strong> (SI)<br />

Teaching Experience<br />

2011-Present OCEE 540 Interpreting our Environment, Nova Southeastern University, Florida<br />

2011-Present OCEE 520 Teaching Environmental Concepts, Nova Southeastern University, Florida<br />

2010-2010 FIW 4314 Conservation Biology, Department of Fisheries & Wildlife Sciences, Virginia Tech<br />

2008-2010 BIOL 1130 General Biology, Department of Biology, Georgia Southern University<br />

2009-2010 BIOL 2108L Biology of Organisms Lab, Department of Biology, Georgia Southern University<br />

2009-2010 BIOL 2108SI Biology of Organisms Supplemental Instruction, Georgia Southern University<br />

2006-2011 Behavior, Ecology, and Conservation of Manatees & Dolphins in Belize, Independent Course with<br />

approved credit hours accepted by ~ a dozen accredited U.S. universities.<br />

2007-2009 EW/NCRC West African Fellows Manatee Conservation Workshop, Ghana, West Africa<br />

2008 Substitute for high school science classes in Spotsylvania, Stafford, & Fredericksburg, VA<br />

2004-2008 Marine Biology for Primary & Secondary Teachers Workshop, Belize Ministry of Education<br />

2004-2008 Biology of Mangroves, Sea Grass & Coral Reefs Field Course, Belize Ministry of Education<br />

1993-1997 Marine Biology, Chemistry, Mathematics, and Scientific Writing tutor, Coastal Carolina University<br />

1986-1993 Principals of Real Estate, Real Estate Brokerage, Real Estate Finance, Fair Housing, and Continuing<br />

Education Courses, Fredericksburg Area Association of Realtors, Virginia Real Estate Commission<br />

Additional Training in Education<br />

2011 KPA: Karen Pryor Academy Certified Training Partner<br />

2011 TTI: TAGteach <strong>International</strong> Certification<br />

2011 CCPDT: Real Solutions for Canine Behavior Problems by Pat Miller (June 11-12, 2011)<br />

2011 NSU: Blackboard Learn Distance-Education Training<br />

Caryn Self-Sullivan CV Page 1 of 1 Updated<br />

12/10/2011


2010 VT: Scholar, BANNER, Safe Zone Training<br />

2010 GSU: Teaching Large Classes, Teaching What You Don’t Know, Thinking Like Leonardo, Safe Space<br />

Training, Creating a Safe Learning Environment; Service-Learning "All for One and One for All<br />

Integrating Service with Teaching and Scholarship for Maximum Impact"; Successful Grant Writing<br />

Workshop; Inertia vs. Critical Mass Engaging Students in Active Discussions, Team-Based Learning-A<br />

Transformational Use of Small Groups in College Teaching; Scholarship of Teaching and Learning<br />

Commons; Teaching Large Classes FLC; Lab Safety Training, VISTA/WebCT/Blackboard/GeorgiaView;<br />

Service-Learning; CPS Classroom System; WINGS, BANNER, The Art of Changing the Brain; Teaching<br />

Large Classes FLC; POGIL Workshop; Active Learning in Large Classes; Active Learning for Students<br />

Graduate Students<br />

2011-Present Kristin Montour-Grubbs, Nova Southeastern University, M.S. Student, Committee Chair<br />

2011-Present Allen Conrad Allen, Nova Southeastern University, M.S. Student, Mentor<br />

2006-Present Haydee Dominguez, Duke University, Ph.D. Student, Mentor, Field Advisor, External Advisor. 2011<br />

Proposal: Manatees in the Dominican Republic.<br />

2006-2008 Marie-Lys Bacchus, Loma Linda University. M.S., Mentor, Field Advisor, Committee Member. 2007<br />

Thesis: Characterization of Resting Holes and Use by the Antillean Manatee (Trichechus manatus<br />

manatus).<br />

2009-2011 Kamla Aristide, University of Dschang, Cameroon, M.S. Student. Mentor, Co-advisor. 2011 Thesis:<br />

Activity Center, Habitat Use and Conservation of the West African Manatee (Trichechus senegalensis<br />

Link, <strong>17</strong>95) in the Douala-Edea Wildlife Reserve and Lake Ossa Wildlife Researve.<br />

2008-2009 Andrea Tanny, Georgia Southern University, M.S. Student, Mentor, Field Advisor, Committee Member.<br />

MS Proposal: 2008 Foraging Behavior in West Indian Manatees.<br />

Publications<br />

Appletons, W. et al. (in review). Magnitude of global marine biodiversity: one third of sea creatures discovered.<br />

Submitted to Science on 15 October 2010.<br />

Bacchus, M-L, S. G. Dunbar, C. Self-Sullivan. 2009. Characterization of resting holes and their use by the Antillean<br />

manatee (Trichechus manatus manatus) in the Drowned Cayes of Belize. Aquatic Mammals 35(1)62-71.<br />

Deutsch, C. J., C. Self-Sullivan, and A. A. Mignucci-Giannoni. 2007. Trichechus manatus. In 2007 IUCN Red List of<br />

Threatened Species (www.iucnredlist.org).<br />

Hoffman, M., et al. 2010. The impact and shortfall of conservation on the status of the world’s vertebrates. Published<br />

in Science Express online 10/26/2010; pending print publication date.<br />

LaCommare, K. S., C. Self-Sullivan, and S. Brault. 2008. Distribution and Habitat Use of Antillean Manatees<br />

(Trichechus manatus manatus) in the Drowned Cays Area of Belize, Central America. Aquatic Mammals<br />

34(1)34-43.<br />

LaCommare, K.S., S. Brault, C. Self-Sullivan, E.M. Hines. (in review). A Boat-based Method for Monitoring<br />

<strong>Sirenian</strong>s: Antillean Manatee Case Study.<br />

Schipper, J., et al. 2008. The status of the world's land and marine mammals: diversity, threat, and knowledge.<br />

Science 322225-230, 10 October 2008. DOI 10.1126/science.1165115.<br />

Self-Sullivan, C. 2004-2010. Manatee and mangrove special topics in the 6th, 7th, & 8th editions of Moon<br />

Handbooks Belize by Joshua Berman and Chicki Mallan, Avalon Travel Publishing, Emeryville, CA.<br />

Self-Sullivan, C. 2008. Conservation of Antillean Manatees in the Drowned Cayes Areas of Belize. Ph.D.<br />

Dissertation, Department of Wildlife and Fisheries Sciences, Texas A&M University.<br />

Self-Sullivan, C. (in press). West Indian Manatees (Trichechus manatus) in the Wider Caribbean Region. Book<br />

Chapter in <strong>Sirenian</strong> Conservation Issues and Strategies in Developing Countries. Edited by Ellen Hines, John<br />

Reynolds, Lemnuel Aragones, Antonio A. Mignucci-Giannoni, Miriam Marmontel. University of Florida<br />

Press.<br />

Self-Sullivan, C. and A. A. Mignucci-Giannoni. 2007. Trichechus manatus manatus. In 2007 IUCN Red List of<br />

Threatened Species (www.iucnredlist.org).<br />

Self-Sullivan, C., and L. Bird. 2007. Marine Science for Students Workbook. BRC Publishers, Benque, Belize.<br />

Self-Sullivan, C., Smith, G. W., Packard, J. M., and LaCommare, K. S. 2003. Seasonal occurrence of male Antillean<br />

manatees (Trichechus manatus manatus) on the Belize Barrier Reef. Aquatic Mammals 29(3)342-354.<br />

Williams, Jr., E. H., Mignucci-Giannoni, A. A., Bunkley-Williams, L., Bonde, R. K., Self-Sullivan, C., Preen, A., and<br />

Cockcroft, V. G. 2003. Echeneid-sirenian associations, with information on sharksucker diet. Journal of Fish<br />

Biology 631<strong>17</strong>6-1183.<br />

Caryn Self-Sullivan CV Page 2 of 2 Updated<br />

12/10/2011


Unpublished Reports<br />

Ofori-Danson, Patrick, Caryn Self-Sullivan, Victor M. Mombu, and Martin A. Yelibora. 2008. Enhancing<br />

Conservation of the West African Manatee in Ghana 2007 Annual Report. Prepared by Nature Conservation<br />

Research Centre (NCRC), P. O. Box KN925, Accra. June 2008.<br />

Self-Sullivan, C. 2006. Report to NCRC on an Insular Population of West African Manatees (Trichechus<br />

senegalensis) in the Upper Afram Arm of Volta Lake.<br />

Self-Sullivan, C., and K. S. LaCommare. 1998-2010. Annual Progress Reports to Belize Forestry Department, Belize<br />

Fisheries Department, Belize Coastal Zone Management Authority & Institute, Friends of Swallow Caye,<br />

Earthwatch Institute.<br />

Self-Sullivan, C., M. Fogel, and M. J. Wooller. 2005. Carbon and nitrogen stable isotope analyses of Antillean<br />

manatees from three distinct ecosystems in Belize inference about relative differences in diets and habitats.<br />

Preliminary Report to Wildlife Trust Belize, 30 October 2005.<br />

Conference Presentations & Posters<br />

Bonde, R. K., P. Lewis, D. A. Samuelson, C. Self-Sullivan, N. E. Auil, and J. A. Powell. 2006. Belize manatee<br />

(Trichechus manatus manatus) epibionts SEM viewing techniques. Poster Presentation at SMM 16 th Biennial<br />

Conference on the Biology of Marine Mammals, San Diego, 12-16 December 2005.<br />

LaCommare, K. S., C. Self-Sullivan, and S. Brault. 2005. Distribution, habitat use and seasonal occurrence of<br />

Antillean manatees in the Drowned Cayes area of Belize, Central America. Oral Presentation at the SMM<br />

16th Biennial Conference on the Biology of Marine Mammals, San Diego, 12-16 December 2005.<br />

LaCommare, K. S., Sullivan, C. S., Brault, S. 2001. Distribution and Foraging Ecology of Antillean Manatees<br />

(Trichechus manatus) in the Drowned Cays Area of Belize, Central America. Poster Presentation at the SMM<br />

14 th Biennial Conference on the Biology of Marine Mammals, Vancouver, British Columbia, 28 November - 3<br />

December 2001.<br />

Mignucci-Giannoni, A. A., and C. Self-Sullivan. 2005. Conservation Status of the Antillean Manatee (Trichechus<br />

manatus manatus) in the Wider Caribbean Region. Oral Presentation at the <strong>International</strong> <strong>Sirenian</strong> Workshop,<br />

SMM 16 th Biennial Conference on the Biology of Marine Mammals, San Diego, 12-16 December 2005.<br />

Self-Sullivan, C. 2009. Conservation of Antillean manatees in Belize 1998-2008. Oral presentation at the 2009<br />

<strong>International</strong> <strong>Sirenian</strong> Conservation Conference, 23-24 March 2009.<br />

Self-Sullivan, C. 2009. Triumph on the commons in Belize the importance of traditional knowledge and stakeholder<br />

input to successful MPAs. Oral Presentation at the <strong>Sirenian</strong> Workshop, Improving the Contribution of Marine<br />

Protected Areas to the Conservation of <strong>Sirenian</strong>s. Society for Conservation Biology Marine Section,<br />

<strong>International</strong> Marine Conservation Congress, May 21-24, 2009, Johnson Center, George Mason University,<br />

Fairfax, Virginia, USA.<br />

Self-Sullivan, C. 2008. Conservation of Antillean manatees in Belize. Seminar (3 hours) at Coastal Zone<br />

Management Authority and Institute, Belize City, Belize.<br />

Self-Sullivan, C. 2006. Non-lethal boat scars on manatees in Belize as a tool for evaluation of a Marine Protected<br />

Area. Oral Presentation at the 59th Annual Gulf and Caribbean Fisheries Institute Conference, Belize City, 6-<br />

11 November 2006 (Science and Management of Marine Protected Areas Session).<br />

Self-Sullivan, C. 2007. Community Conservation of West African Manatees in the Upper Afram Arm of Volta Lake,<br />

Ghana. Oral Presentation, Marine Mammals & Indigenous Communities Workshop, SMM <strong>17</strong>th Biennial<br />

Conference on the Biology of Marine Mammals, Cape Town, November 2007.<br />

Self-Sullivan, C. 2007. Conservation and Status of West African Manatees in Ghana, Senegal, Nigeria, Cameroon,<br />

and Sierra Leone. Oral Presentation, Marine Mammals in Africa Workshop, SMM <strong>17</strong>th Biennial Conference<br />

on the Biology of Marine Mammals, Cape Town, November 2007.<br />

Self-Sullivan, C. LaCommare, K.S. 2003. One Issue Related to Research and Conservation: Cruise Ship Tourism in<br />

Belize. Oral Presentation at the <strong>International</strong> <strong>Sirenian</strong> Workshop, SMM 15 th Biennial Conference on the<br />

Biology of Marine Mammals, Greensboro, North Carolina, 14 December 2003.<br />

Self-Sullivan, C., and A. A. Mignucci-Giannoni. 2006. Conservation Status of the Antillean Manatee (Trichechus<br />

manatus manatus) in the Wider Caribbean Region. Oral Presentation at IMC9, Sapporo, Japan, 31 July 2005.<br />

Self-Sullivan, Caryn. 1996. A Survey of Macroalgae in Discovery Bay, Jamaica Species Diversity and Zonation<br />

Differences in the Lagoon. Senior Thesis, Coastal Carolina University, SC Oral presentation to Discovery<br />

Bay Lab, Jamaica.<br />

Self-Sullivan, Caryn. 1996. The Effects of Benthos on Acartia tonsa Population Dynamics. National Conference on<br />

Undergraduate Research, Ashville, NC POSTER - Results of NSF-REU Fellowship research. Oral<br />

Caryn Self-Sullivan CV Page 3 of 3 Updated<br />

12/10/2011


presentation to Horn Point Environmental Lab and Coastal Carolina University.<br />

Sullivan, C. S., LaCommare, K. S., Packard, J. M., and Evans, W. E. 2001. Seasonal Occurrence of Male Antillean<br />

Manatees (Trichechus manatus manatus) on the Belize Barrier Reef and the Effect of Scale on Seasonal<br />

Distribution Studies. Poster Presentation at the SMM 14 th Biennial Conference on the Biology of Marine<br />

Mammals, Vancouver, British Columbia, 28 November - 3 December 2001.<br />

Sullivan, C. S., Packard, J. M., and Evans, W. E. 1999. Spring distribution and behavior of Antillean manatees in the<br />

Drowned Cayes, Belize. Poster Presentation at the SMM 13th Biennial Conference on the Biology of Marine<br />

Mammals, Maui, Hawaii, November 28 - December 3, 1999.<br />

Achievements, Awards, Honors<br />

• 2011 B.F. Skinner Scholarship, KPA<br />

• 2009 CCU COS Alumnus of the Year<br />

• 2003 Earthwatch Young Scientist Award<br />

• 1998-2001 NSF Graduate Fellowship<br />

• 1997 Outstanding Marine Science Senior<br />

• 1997 President's Excellence Nominee<br />

• 1996 Exchange student to Deakin<br />

University<br />

• 1995 NSF-REU Fellowship<br />

• 1994-1997 Nelson Scholarship<br />

• 1993 Freshman Essay Contest<br />

• 1993-1997 President's Honor List<br />

• 1993-1997 Dean's Honor List<br />

• Sigma Zeta Beta Mu Honor Society<br />

• Omicron Delta Kappa Society<br />

• Pi Mu Epsilon Honor Society<br />

Grants and Fellowships<br />

• Marine Mammal Commission 2009. <strong>Sirenian</strong>/MPA Workshop. Amount funded $12,105.<br />

• Earthwatch Institute, Center for Research 2001-2007. Field Research. Amount funded ~ $56,000/year<br />

• UNDP/GEF/COMPACT, 2006. HPF Educational Outreach Programme. Amount funded $40,000USD<br />

• Project AWARE Small Grant Program 2004. Field Research. Amount Funded $1,000<br />

• Earthwatch Institute, Center for Research 2003 Young Scientist Award. Amount funded $2,500<br />

• Conservation Action Fund, New England Aquarium 2001. Capacity Building Grant. Amount funded $5,000<br />

• American Museum of Natural History Lerner-Grey Fund for Marine Research 2000. Field Research. Amount<br />

funded $2,000.<br />

• Oceanic Society 1998-1999. Field Research. Amount funded undisclosed (9 months of field research in Belize)<br />

• National Science Foundation Graduate Fellowship 1998-2001. Tuition & Stipend. Amount funded $75,000<br />

Other Service Projects<br />

• Workshop Organizer Society for Conservation Biology, 2009 IMCC Conference, <strong>Sirenian</strong>s and MPAs<br />

• Peer reviewer for academic journals: Aquatic Mammals, Behaviour, Journal of Wildlife Management, USGS<br />

Sirenia Project, Journal of Environmental Studies and Sciences, Estuaries & Coasts<br />

• Scientific Advisor to field scientists from Guatemala, Ghana, and Dominican Republic regarding techniques for<br />

evaluating manatee populations in remote areas, with an emphasis on the incorporation of manatee behavior<br />

• Mentor young people (especially women) entering the sciences (M.S., Ph.D. students and interns)<br />

• Science Judge National Oceans Science Bowl Regional Competition, Texas A&M, 2006<br />

• Coordinator National Marine Science Symposium, Belize City, Belize 2006<br />

• Coordinator National Marine Science Quiz Bowl, Belize City, Belize 2006-2007.<br />

• Committee Member Society for Marine Mammalogy Conference Committee, 2003<br />

Educational Outreach Publications and Presentation: >100 public presentations on animal behavior, marine mammals,<br />

marine and environmental science through NGO and academic organizations including: public schools & dive clubs, the<br />

Sierra Club, Green Drinks, New England Aquarium, Coastal Carolina University, Mary Washington University, Texas<br />

A&M University, Universidad Autonoma de Santo Domingo, Rutgers University, Georgian Court University, Brookdale<br />

Community College, The Wetlands Institute, and at community festivals, etc.<br />

References<br />

Jane Dougan, Distance-Ed Coordinator, Nova Southeastern University, douganj@nova.edu, 866.229.2557<br />

Stephen P. Vives, Ph.D., Georgia Southern University, Statesboro, GA, svives@georgiasouthern.edu, 912.478.5487<br />

Bruce A. Schulte, Ph.D., Western Kentucky University, Bowling Green, KY, bruce.schulte@wku.edu, 270.745.5999<br />

Jessica R. Young, Ph.D., Western State College of Colorado, jyoung@western.edu, 970.943.3045<br />

Michelle Kopfer, Teaching Assistant, Virginia Tech, Blacksburg, VA, midavis1@vt.edu, 540.577.1081<br />

Paul J. Camp, Ph.D., Professor, Spelman College, pjcamp@gmail.com, 404.270.5864<br />

Jane M. Packard, Ph.D., Professor, Texas A&M University, College Station, TX, j-packard@tamu.edu, 979.845.1465<br />

Wyndylyn von Zharen, Ed.D., J.D., Professor, Texas A&M University, dr_vonzharen@msn.com, 409.740.4485<br />

Caryn Self-Sullivan CV Page 4 of 4 Updated<br />

12/10/2011


145 S. Tompkins Street<br />

Howell, MI 48843<br />

K ATHERINE S PENCER L A C OMMARE<br />

EDUCATION<br />

PhD. June 2011. Environmental Biology, University of Massachusetts, Boston, MA<br />

MSc. 1994. Forestry and Natural Resources, Purdue University, West Lafayette, IN<br />

B.Sc. 1989. Anthropology/Zoology, University of Michigan, Ann Arbor, MI<br />

Kslacommare@gmail.com<br />

(248) 756-3985<br />

TEACHING EXPERIENCE<br />

Instructor: Introduction to Environmental Science (2004, 2008-2011). Face-to-face; Hybrid,<br />

Online. Lansing Community College, Lansing, MI,<br />

Instructor: Organismal Biology (2010-2011). Face-to-face. Lansing Community College, Lansing,<br />

MI,<br />

Instructor: Behavior and Conservation of Marine Mammals Field Course. (2008-2010) Spanish<br />

Bay Research and Education Center, Belize, C.A.<br />

Field Instructor: Coastal Ecology and Marine Mammal Ecology (1998-2007) Oceanic Society,<br />

Earthwatch<br />

Teaching Assistant: Cell Biology (2001). University of Massachusetts, Boston<br />

Teaching Assistant: General Biology (2001). University of Massachusetts, Boston<br />

Writing Instructor: Introduction to Environmental Science (1999-2000), University of<br />

Massachusetts, Boston,<br />

PROFESSIONAL DEVELOPMENT AND CERTIFICATIONS<br />

University of Michigan Community College Biology Active Learning Workshop: August<br />

2011, Ann Arbor, MI<br />

Teaching Ecology to Undergraduates: Spring 2011. Pierce Creek Institute, Hastings MI.<br />

What should college biology look like? the first two years: Spring 2001. Michigan State<br />

University<br />

Leading Effective Discussions: Spring 2010. Lilly Seminar Series <strong>–</strong> Michigan State University<br />

Transforming Learning Through Teaching: Spring 2009. 14-week professional development<br />

course through Lansing Community College’s Center for Teaching Excellence.<br />

Michigan Virtual University Certification: Fall 2008<br />

PROFESSIONAL EXPERIENCE<br />

Principal Investigator: January 2001 <strong>–</strong> December 2003, Co-PI: January 2004 <strong>–</strong> Dec. 2007, Manatees in<br />

Belize, Earthwatch Project (part-time, contract)<br />

• Managed all project logistics: developed and managed the budget, hired and supervised staff,<br />

recruited and hired student interns, trained interns and volunteers, completed endangered species<br />

permit applications, wrote agency reports, wrote and received grants, hired vendors, conducted<br />

government relations and managed volunteer activities and developed and managed education<br />

activities.


• Developed and managed educational activities including: determining priorities, developing<br />

curriculum and leading expedition activities for students and volunteers.<br />

• Designed research protocol. Collected, analyzed, presented and published project data. Utilized<br />

Excel, SPSS, MatLab.<br />

• Managed data collection, training, education and field trips for 1-5 two-week teams per year of 8-24<br />

volunteers per team.<br />

• Managed project logistics: budgets ($43,000 per year), permits, agency reports, and vendors.<br />

• Managed 2-3 field assistants and student interns per team.<br />

• Networked with local agencies and NGO’s to develop an appropriate manatee research program.<br />

• Conducted 1-3 public/professional speaking events per year.<br />

• Authored or co-authored 4 scientific papers.<br />

• Designed and maintained an Access database to archive, organize and house individual manatee,<br />

manatee population, habitat and seagrass data.<br />

Aerial observer, (Part-time - contract) April - May 2001, National Marine Fisheries Service, Northeast<br />

Fisheries Science Center, Woods Hole, MA. Supervisor: Pat Gerrior.<br />

• Participated in line transect surveys from a twin engine Otter for right and other large whales off<br />

the Massachusetts coast.<br />

Fisheries Research Biologist, May 1999 <strong>–</strong> August 2000, National Marine Fisheries Service, Northeast<br />

Fisheries Science Center. CMER Assistantship program, Woods Hole, MA. Supervisor: Tim Smith<br />

• Analyzed distribution and habitat ecology of Brydes whales in the North Pacific with ArcView GIS<br />

and SPlus statistical package.<br />

Field Naturalist and Biologist: (Contract Position) October 1998 <strong>–</strong> March 2000, Oceanic Society<br />

Expeditions, Belize Dolphin/Manatee Project, Drowned Cayes, Belize, Central America<br />

• Designed research methods, collected and analyzed data on the habitat and foraging ecology of<br />

manatees.<br />

• Coordinated volunteer data collection.<br />

• Coordinated daily educational and natural history activities for volunteers.<br />

Land Protection Specialist: March 1995 <strong>–</strong> October 1998, The Nature Conservancy, Michigan Field<br />

Office, East Lansing, MI, Supervisor: Bill McCort<br />

• Developed and managed a landowner education program to promote conservation stewardship of<br />

private lands. Program activities included educating landowners about endangered habitats and<br />

assisting them in maintaining these habitats on their property.<br />

Natural Areas Preservation Assistant: October 1994 <strong>–</strong> March 1995, Ann Arbor Parks and Recreation<br />

Department, Ann Arbor, MI, Supervisor: David Borneman<br />

• Assisted Natural Areas Preservation Coordinator with general duties.<br />

• Managed plant inventory database.<br />

• Coordinated volunteer workdays in city parks.<br />

PUBLICATIONS<br />

LaCommare, Katherine et al. (in review). A Boat-based Method for monitoring <strong>Sirenian</strong>s: Antillean<br />

Manatee Case Study. Biological Conservation<br />

Aragones, Lemenuel; Katherine S. LaCommare, Sarita Kendall, Delma Nataly Castelblanco-<br />

Martinez, Daniel Gonzalez-Socoloske. In press. Boat and Land-based Surveys for <strong>Sirenian</strong>s in<br />

Developing Nations. In Hines et al. <strong>International</strong> Strategies for Manatee and Dugong<br />

Conservation. Florida University Press<br />

LaCommare, Katherine et al. 2008. Distribution and Habitat Use of Antillean Manatees in the Drowned<br />

Cayes Area of Belize, Central America. Aquatic Mammals Journal. 34(1): 35-42


Self Sullivan, Caryn, Gregory W. Smith, Jane M. Packard, Katherine S. LaCommare. 2003.<br />

Seasonal Occurrence of Male Antillean Manatees (Trichechus manatus manatus) on the Belize Barrier Reef.<br />

Aquatic Mammals. 29(3): 342-354.<br />

Weiler, Katherine S. and Joseph T. O’Leary. May 1994. Demographic Change and Forest Resources:<br />

Implications for the Lake States. Lake States Forest Resource Assessment.<br />

PUBLICATIONS IN PREPARATION<br />

LaCommare, Katherine et al. (in prep). Coastal Development in Belize: Impact on Manatees and<br />

Seagrass.. Aquatic Botany.<br />

LaCommare, Katherine et al. (in prep). Conservation Planning for Antillean Manatees in Belize.<br />

Conservation Biology.<br />

POPULAR PUBLICATIONS<br />

LaCommare, Katherine S. Fall 2000. <strong>Sirenian</strong> Song. Fall newsletter for the Evergreen Group’s Manatee<br />

ambassador program<br />

LaCommare, Katherine S. Spring 2000. <strong>Sirenian</strong> Song. Spring newsletter for the Evergreen Group’s<br />

Manatee ambassador program<br />

LaCommare, Katherine S. Fall 1999. <strong>Sirenian</strong> Song. Fall newsletter for the Evergreen Group’s Manatee<br />

ambassador program<br />

Adimey, Nicole M. and Katherine S. Weiler. October 1994. Orcas of Johnstone Strait. Waterway Times<br />

Weiler, Katherine. December 1994. The Bald Eagle Soars, But What About the Fuzzy Sandoze? The Bard<br />

PRESENTATIONS AT NATIONAL CONFERENCES<br />

LaCommare, Katherine et al May 2011. Oral Presentation. A Boat-based method for monitoring<br />

<strong>Sirenian</strong>s: Antillean Manatee Case Study. Oral Presentation. <strong>International</strong> Marine Conservation<br />

Congress. Victoria, BC Canada. May 2011.<br />

LaCommare, Katherine et al. May 2009. Mangrove Clearing and Bottom Dredging: How is Antillean<br />

Manatee (Trichechus manatus manatus) Habitat Use Affected by Coastal Development in Belize?. Oral<br />

Presentation. <strong>International</strong> Marine Conservation Congress. Washington, DC. May 2009.<br />

LaCommare, Katherine et al. December 2005. Distribution, Seasonal Occurrence and Habitat Use of<br />

Antillean Manatees in the Drowned Cayes Area of Belize, Central America. Oral Presentation. 16th Biennial Conference on the biology of Marine Mammals. San Diego, California. November<br />

2005.<br />

LaCommare, Katherine et al. December 2003. Habitat Use of Seagrass in the Drowned Cayes Area of<br />

Belize, Central America. Poster Presentation. 15th Biennial Conference on the biology of Marine<br />

Mammals. Greensborough, SC.<br />

LaCommare, Katherine et al. December 2001. Habitat Use of Antillean Manatees in the Drowned Cayes<br />

Area of Belize, Central America. Poster Presentation. 14th Biennial Conference on the biology of<br />

Marine Mammals. Vancouver, BC<br />

INVITED PRESENTATIONS<br />

LaCommare, Katherine S. April 2002. Habitat Ecology of Antillean Manatees in the Drowned Cayes Area<br />

of Belize. Earthwatch Chicago, Spring Event, Kent Culinary Institute, Chicago, IL<br />

LaCommare, Katherine S. March 2002. Habitat Ecology of Antillean Manatees in the Drowned Cayes<br />

Area of Belize. Science in Action Lecture Series, Maritime Aquarium, Norwalk, CT


LaCommare, Katherine S. November 2001. Habitat Ecology of Antillean Manatees in the Drowned Cayes<br />

Area of Belize. Lowell Lecture Series, New England Aquarium, Boston, MA<br />

FUNDING AND AWARDS<br />

June 2007 Dissertation Improvement Grant, University of Massachusetts, Office of Sponsored<br />

Programs. Project Title: Ecology and Behavior of Antillean Manatees in the Drowned Cays, Belize, Central<br />

America. Funding: $1,200<br />

January 2003 <strong>–</strong> December 2006. Co-Principal Investigator. Earthwatch Institute. Project Title: Ecology<br />

and Behavior of Antillean Manatees in the Drowned Cays, Belize, Central America. Funding: $43,000/year<br />

October 2002. Young Scientist Award. Earthwatch Institute. Project Title: Ecology and Behavior of<br />

Antillean Manatees in the Drowned Cays, Belize, Central America. Funding: $5,000<br />

January 2001 <strong>–</strong> December 2003. Principal Investigator. Project Title: Ecology and Behavior of Antillean<br />

Manatees in the Drowned Cays, Belize, Central America. Funding: $43,000/year<br />

December 2000. Conservation Action Fund. Project Title: Ecology and Behavior of Antillean Manatees in the<br />

Drowned Cays, Belize, Central America. Award: $5,120<br />

ADVISORY BOARDS/PROFESSIONAL SOCIETIES<br />

Livingston Land Conservancy, Board of Directors, Fundraising and Membership Chair, (2010present)<br />

Livingston Land Conservancy, Fundraising Committee (2004, 2010)<br />

<strong>Sirenian</strong> <strong>International</strong>, Inc., Co-founder<br />

<strong>Sirenian</strong> <strong>International</strong>, Inc.,Board of Directors, Secretary, (2000-2005)<br />

Michigan Science Teachers Association, Member<br />

National Science Teachers Association, Member<br />

Michigan Association of Community College Biology Teachers, Member<br />

Science Department Curriculum Committee, Adjunct Faculty Representative<br />

REFERENCES<br />

Meg Clark Elias, Professor, Science Department, Lansing Community College, Lansing, MI 48933. (5<strong>17</strong>)<br />

483-1557, Clarkm1@lcc.edu<br />

Solange Brault, Associate Professor, University of Massachusetts, Boston, Department of Biology, 100<br />

Morrissey Blvd., Boston, MA 02125. (6<strong>17</strong>) 287-6683, Solange.brault@umb.edu<br />

Robert Stevenson, Associate Professor, University of Massachusetts, Boston, Department of Biology, 100<br />

Morrissey Blvd., Boston, MA 02125. (6<strong>17</strong>) 287-6579, Robert.stevenson@umb.edu<br />

Caryn Self Sullivan, Adjunct Faculty, Nova Southeastern University, President, <strong>Sirenian</strong> <strong>International</strong>. (540)<br />

287-8207, cselfsullivan@sirenian.org<br />

William D. McCort, Ph.D., Director and Manager of Canine Professor LLC, Chair, Natural Areas Technical<br />

Advisory Committee, Natural Areas Preservation Program, Washtenaw County Parks & Recreation<br />

Department, Volunteer, Land Protection Committee, Washtenaw Land Trust, 18819 Bush Road,<br />

Chelsea, MI 48118-9749. 734-475-6908, billmccort@comcast.net


CURRICULUM VITAE<br />

HEATHER J. KALB<br />

Department of Natural Sciences and Mathematics, West Liberty University,<br />

West Liberty University, West Liberty, WV 26074<br />

(812) 461-8965 (C); heather.kalb@westliberty.edu<br />

Education<br />

Ph.D. in Zoology, 1999. Dept. Biology, Texas A&M University, College Station, TX. “Behavior and<br />

Physiology of Solitary and Arribada Nesting Olive Ridley Sea Turtles during the Internesting<br />

Period.” General areas of specialization: chelonians, conservation, reproductive physiology and<br />

behavior. Academic Advisor: Dr. David W. Owens<br />

B.A., 1989. Major Biology, Minor Chemistry, Wittenberg University, Springfield, OH.<br />

Positions Held<br />

Assistant Professor, Dept. of Natural Sciences and Mathematics, West Liberty University, Wheeling,<br />

WV 26074. Aug. 2010 - current<br />

Assistant Professor (temp.), Dept. of Biology, Georgia Southern University, Statesboro, GA 30460.<br />

Aug. 2008 - 2010<br />

Assistant Professor, Dept. of Biology, University of Evansville, Evansville, IN 47722.<br />

Aug. 2003 - May 2008<br />

Assistant Professor (temp.), Dept. of Biology, Arkansas State Univ., State University, AR 72467.<br />

Aug. 2002 - May 2003<br />

Lecturer, Dept. of Biology, West Chester University, West Chester, PA 19383.<br />

Aug. 2000-May 2002<br />

Teaching Assistant, Dept. of Biology, Texas A&M University, College Station, TX 77845.<br />

1989-92, 1994-99. Responsibilities: teach labs, write/grade quizzes & exams, hold office hours.<br />

Course Director / Assistant Lecturer for “Advanced Sea Turtle Biology” course. Jan. 1998.<br />

Méxican Sea Turtle Center in Mazunte, Oaxaca, México.<br />

Primary instructors: D. Owens (Dept. of Biology, Texas A. & M. Univ.), J. Flanagan<br />

(Veterinarian, Houston Zoo), R. Marquez (National Coordinator, Sea Turtle Research, Méxican<br />

National Fisheries Institute)<br />

Students: 15 Méxican/Cuban government researchers, veterinarians, and students.<br />

Range of topics: reproductive physiology & behavior, disease, and life history and hands-on<br />

training with laparoscopy, ultrasonography, blood sampling, and radio-telemetry.<br />

Assistant Lecturer/Organizer for “Reproductive Physiology of Sea Turtles” course.<br />

2-week course at Rancho Nuevo, México. May 1996.<br />

Courses Taught (locations where the course was taught are in parenthesis)<br />

Intro level courses, Majors<br />

• Animal Diversity/Introductory Zoology, lecture and lab (UE, ASU, WCU)<br />

• Biology 1 lab (GSU, UE, TAMU)<br />

• Biology 2 lab (GSU, TAMU)<br />

• Scientific Communications (UE)<br />

Upper level courses, Majors<br />

• Comparative Animal Physiology lecture and lab (GSU, UE, WLU)<br />

1


• Comparative Vertebrate Anatomy, lecture and lab (WCU, TAMU)<br />

• Animal Behavior, lecture (UE, WLU)<br />

• Vertebrate Zoology, lecture and lab (UE, WLU)<br />

• Conservation Biology, lecture (WCU)<br />

• Vertebrate Ecology, lecture and lab (WCU)<br />

• Undergraduate research (individual student) (GSU, UE)<br />

Biology courses for (usually) non-majors<br />

• Human Anatomy and Physiology 1, lecture and lab (GSU, WCU)<br />

• Human Anatomy & Physiology II, lecture and lab (WCU)<br />

• Microbiology lecture (summer 2010) (GSU)<br />

• Microbiology lab (scheduled for fall 2010) (GSU)<br />

• General Biology, lecture and lab (GSU, UE, ASU, WLU)<br />

• Environmental Science, lecture (UE)<br />

Honors, Fellowships, Other Recognitions<br />

Affiliate status for the Graduate Faculty at Georgia Southern University. 2010.<br />

Clare Booth Luce Fellowship in Science and Engineering ($25,000/yr). 1992-1994.<br />

Outstanding Teaching Performance, Dept. of Biology, Texas A&M University. 1992.<br />

Alpha Lambda Delta, Honor Society. 1985.<br />

Wittenberg University Scholar (top 5% of incoming freshman). 1985-1989.<br />

Reviewed Book Chapter, Peer Reviewed Publications, and Professional Reports<br />

S. Morreale, P. Plotkin, D. Shaver, and H. Kalb. 2007. Adult Migration and Habitat Utilization. In<br />

Biology and Conservation of Ridley Turtles. P. Plotkin and S. Morreale, Editors. John Hopkins<br />

University Press.<br />

G. Zug, H. Kalb, S. Lazar. 1997. Age and growth in wild Kemp’s ridley sea turtles (Lepidochelys<br />

kempii) from skeletochronological data. Biological Conservation 80: 261-268.<br />

H. Kalb and G. Zug. 1990. Age estimates for a population of American toads, Bufo americanus<br />

(Salientia: Bufonidae), in northern Virginia. Brimleyana 16:79-86.<br />

H. Kalb. 1992. Reproductive Physiology of the Spotted Turtle, Clemmys guttata at the Prairie Road<br />

Fen, in Springfield Ohio. Final Report to the Ohio Dept. of Natural Resources.<br />

H. Kalb and L. Laux. 1990. A Population Study of Clemmys guttata at the Prairie Road Fen, in<br />

Springfield Ohio. Final Report to the Ohio Dept. of Natural Resources.<br />

Editor<br />

H. Kalb, A. Rohde*, K. Gayheart, and K. Shanker. 2008. Proceedings of the 25th Annual<br />

Symposium on Sea Turtle Biology and Conservation. NOAA Tech. Memo.<br />

H. Kalb and T. Wibbels. 2000. Proceedings of the 19th Annual Symposium on Sea Turtle Biology<br />

and Conservation. NOAA Tech. Memo. NMFS-SEFSC-443. pp. 291.<br />

H. Kalb. Turtle and Tortoise Newsletter. 2000-2005. Official newsletter of the IUCN/SSC<br />

Freshwater Turtle and Tortoise Specialist Group. Published by Chelonian Research Foundation.<br />

1 st Edition, Feb. 2000. Editorial Board: Anders Rhodin, Peter C.H. Pritchard, John Behler, and<br />

Russell Mittermeier.<br />

H. Kalb. 1994-1999. Box Turtle Research and Conservation Newsletter. Initial funding: Savannah<br />

River Ecology Lab, later: Chelonian Research Foundation.<br />

*Names in bold are undergraduate students who were working directly with me.<br />

2


Grants and Internships<br />

Faculty Development Grant for travel to the Society for Integrative and Comparative Biology’s<br />

Annual Symposium in Seattle, WA. $1024. January 2010.<br />

The Effects of Incubation Temperatures on Embryonic Development and Post-hatching Morphology<br />

and Behavior in Malayan Box Turtles (Cuora amboinensis). Summer 2006. Funding by<br />

University of Evansville Undergraduate Research Program and Honor’s Program. Students:<br />

Alison Rohde, Katie Aldred, and Natalie Byars.<br />

Management of the Cuora amboinensis Taxon Management Group through Genetic Identification<br />

of Captive Stock. Spring 2006. $1,735. University of Evansville’s Undergraduate Research<br />

Program. Student: Sasha Rohde.<br />

Continued Support for the University of Evansville Captive Breeding Colonies of Asian Turtles.<br />

Spring 2005. $1,359. Univ. of Evansville Alumni Research and Scholarly Activity Fellowship.<br />

Subspecies Identification of Individuals in a Captive Colony of Malayan Box Turtles, Cuora<br />

amboinensis. Summer 2005. Funding Univ. of Evansville Honor’s Program. Student: Andrea<br />

Eyler.<br />

Treatment of Parasites in Malayan Box Turtles, Cuora amboinensis. Summer 2005. $9,810.<br />

Funding Univ. of Evansville Undergraduate Research Program. Student: Bozorgmehr Ouranos.<br />

Ultrasound Technology in the Research Lab and Classroom. 2004. $2,500. U.E. Arts, Research and<br />

Teaching Projects Grant.<br />

Maintenance and Reproduction in Two Asian Turtle Assurance Colonies. Fall 2003, Spring 2004.<br />

Total $5,330. University of Evansville’s Alumni Research and Scholarly Activity Fellowship.<br />

Behavior of the Olive Ridley in the Reproductive Patch. 1993. National Geographic Society. David<br />

Owens grantee.<br />

Reproductive Physiology of Clemmys guttata (Spotted Turtle). Grantee and Primary Investigator.<br />

1991-1992. Ohio Dept. of Natural Resources, Division of Parks and Recreation.<br />

Life History Study of Clemmys guttata. Primary Investigator. 1989.<br />

Ohio Dept. of Natural Resources, Division of Parks and Recreation.<br />

Vertebrate Zoology Internship, Smithsonian Institution, Washington, D.C. 1988. Project:<br />

Skeletochronological Age Estimates on the American Toad and the Kemp’s Ridley Sea Turtle.<br />

Travel Grants (2 @ $500) from Dept. of Biology, Texas A&M Univ. 1995 and 1997. To attend the<br />

77th Annual Meeting of Herpetologists’ League/SSAR and 15 th Annual Sea Turtle Symposium.<br />

Travel Grant ($500) to 2 nd <strong>International</strong> Congress on Chelonian Conservation. From Congress<br />

organizers. 1995.<br />

Research Presentations<br />

T. Hayes and H. Kalb. 2010. Visual Stages of Egg Development in the Malayan box turtle (Cuora<br />

amboinensis) as Observed with Ultrasound Technology. Annual Conference of the Society for<br />

Integrative and Comparative Biology, Seattle WA. (Poster presentation)<br />

N. Byars, A. Rohde, K. Aldred, and H. Kalb. 2007. Effects of Incubation Temperature on<br />

Incubation Duration and Hatchling Fitness in Malayan Box Turtles (Cuora amboinensis). 21 st<br />

National Conference on Undergraduate Research, California. Apr. 12-14. (Poster presentation)<br />

H. Kalb, N. Byars, and K. Aldred. 2006. Observations on Reproduction in a Captive Colony of<br />

Malayan Box Turtles, Cuora amboinensis. 4 th Annual Symposium on Conservation and Biology<br />

of Freshwater Turtles and Tortoises. St. Louis, Mo. Aug 10-13. (Poster presentation)<br />

A. Rohde, B. Ouranos, S. Patton and H. Kalb. 2006. Parasites Present in a Stable Breeding<br />

Population of Malayan Box Turtles (Cuora amboinensis). 4 th Annual Symposium on<br />

Conservation and Biology of Freshwater Turtles and Tortoises. St. Louis, Mo. Aug 10-13.<br />

(Poster presentation)<br />

3


B. Ouranos, C. Bursey, and H. Kalb. 2005. Preliminary Parasite Results in a Captive Colony of<br />

Malayan Box Turtles (Cuora amboinensis). Univ. of Evansville Fall Research Symposium.<br />

(Poster presentation)<br />

A. Eyler, A. Rohde, Y. Hsiang-Jui, and H. Kalb. 2005. Obtaining Blood Samples and Sequence<br />

from Cuora amboinensis: Preliminary Studies for Identifying Subspecies of a Local Collection.<br />

University of Evansville Fall Research Symposium. (Poster presentation)<br />

B. Hart, J. Legout, H. Kalb and M. Davis. 2005. Identification and Antibiotic Sensitivity Testing of<br />

Bacteria Infecting Threatened Asian Freshwater Turtles. 19 th National Conference on<br />

Undergraduate Research, Lexington, VA. Apr. 21-23. (Poster presentation)<br />

H. Kalb. 2005. Preliminary Data on Reproduction in Malayan Box Turtles (Cuora amboinensis). 25 th<br />

Annual Sea Turtle Symposium, Savannah GA. (Oral presentation)<br />

H. Kalb. 2004. Female Reproductive Physiology: a Review, Techniques, and Applications. Turtle<br />

Survival Alliance Symposium, Orlando FL. (Oral presentation)<br />

H. Kalb and D. Owens. 1997. The Olive Ridley Aggregation at Playa Nancite, Costa Rica. 77 th<br />

Annual Meeting of Herpetologists’ League/Society for the Study of Amphibian and Reptiles.<br />

Seattle, WA. (Oral presentation)<br />

H. Kalb and D. Owens. 1995. Internesting Movements and Conservation Concerns of the Olive<br />

Ridley Sea Turtle during the Reproductive Season at Nancite, Costa Rica. <strong>International</strong> Congress<br />

of Chelonian Conservation, Gonfaron, France. (Oral presentation)<br />

H. Kalb and D. Owens. 1995. Nancite to Ostional: Post-mating Movements of Female Olive Ridley<br />

Sea Turtles. 15th Annual Sea Turtle Symposium, Hilton Head, SC. (Oral presentation)<br />

H. Kalb, J. Kureen, P. Mayor, J. Peskin, and R. Phyliky. 1995. Conservation Concerns for the<br />

Nancite Olive Ridleys. 15 th Annual Sea Turtle Symposium, Hilton Head, SC. NOAA Tech.<br />

Memo. NMFS-SEFSC-387:141-142. (Poster presentation)<br />

H. Kalb. 1994. Reproductive Evaluation of Clemmys guttata, the Spotted Turtle. 74 th Annual<br />

Meeting of Herpetologists’ League/Society for the Study of Amphibian and Reptiles. Athens,<br />

GA. (Oral presentation)<br />

H. Kalb and D. Owens. 1994. Differences between Solitary and Arribada Nesting Olive Ridley<br />

Females during the Internesting Period. 14 th Annual Sea Turtle Symposium, Hilton Head, SC.<br />

NOAA Tech. Memo. NMFS-SEFSC-351:68. (Oral presentation)<br />

H. Kalb and D. Owens.1993. Arribada versus Solitary Nesters: Comparison of Internesting<br />

Behavior Patterns in Olive Ridley Sea Turtles, Lepidochelys olivacea. 2 nd World Congress of<br />

Herpetology, Adelaide, Australia. (Oral presentation)<br />

H. Kalb, R. Valverde, and D. Owens. 1992. What is the Reproductive Patch of the Olive Ridley Sea<br />

Turtle at Playa Nancite, Costa Rica? 12 th Annual Sea Turtle Symposium, Jekyll Island, GA.<br />

NOAA Tech. Memo. NMFS-SEFSC-361:57-60. (Oral presentation)<br />

L. Laux, C. Stanton, and H. Kalb. 1991. Spatial Relationship among a Restricted Spotted Turtle,<br />

Clemmys guttata, Population in a Small Prairie Fen. Ecological Society of America Annual<br />

Meeting, San Antonio, TX. (Poster presentation)<br />

D. Rostal, H. Kalb, J. Grumbles, P. Plotkin, and D. Owens. 1991. Application of Ultrasonography to<br />

Sea Turtle Reproduction. 11 th Annual Sea Turtle Symposium, Jekyll Island, GA. NOAA Tech.<br />

Memo. NMFS-SEFSC-302:181. (Poster presentation)<br />

H. Kalb. 1990. A Population Study of Clemmys guttata at the Prairie Road Fen, in Springfield,<br />

Ohio. 70 th Annual Meeting of Herpetologists’ League/SSAR, New Orleans, LA. (Poster<br />

presentation)<br />

G. Zug and H. Kalb. 1989. Skeletochronological Age Estimates for Juvenile Lepidochelys kempii<br />

from Atlantic Coast of North America. 9 th Annual Sea Turtle Symposium Jekyll Island, GA.<br />

NOAA Tech. Memo. NMFS-SEFSC-232:271-273. (Poster presentation)<br />

4


Invited Speaker<br />

The Asian Turtle Crisis and the UE Turtle Program. 2005. Hoosier Herpetological Society.<br />

Indianapolis, IN.<br />

The Asian Turtle Crisis. 2004. University of Southern Indiana, Dept. Biology Spring Seminar Series.<br />

Turtles. 2003. Women in Technology and Science Workshop for 6 th grade girls. Jonesboro, AR.<br />

Olive Ridley Sea Turtles of Playa Nancite, Costa Rica. 2001. Richard Stockton College Biology<br />

Seminar Series<br />

Olive Ridley Sea Turtles of Playa Nancite, Costa Rica. 2001. New York Turtle & Tortoise Seminar.<br />

Box Turtles in the Pet Trade and Sea Turtle Natural History and Conservation. 1997. Dept. of<br />

Recreational Sports, 1 st Annual Jamboree, T.A.M.U, College Station, TX.<br />

Box Turtles: Past, Present, and Future. 1997. Rio Brazos Audubon Society.<br />

Costa Rican Turtle Research & Conservation Implications. 1995. Society for Conservation Biology.<br />

Sea Turtle Natural History & Conservation. 1991. Texas Environmental Action Coalition<br />

Conference.<br />

Scientific American Frontiers TV show. 1991. 10 min. on my chemosensory imprinting research in<br />

Kemp’s ridleys.<br />

University Service Work<br />

University of Evansville, committee work<br />

Interdisciplinary Studies Committee, Chair<br />

Admissions and Standards Committee<br />

Biology library liaison<br />

Phi Eta Sigma (freshman honor’s society) advisor<br />

University of Evansville Turtle Program: Tours and Turtle Conservation Talks<br />

Fall, 2004 & 2005: Scott School.<br />

Spring, 2005: Memorial High School Environmental Science Classes. Contact Maryann Watson.<br />

Spring, 2004 & 2005: Luce Elementary School. Contact Pat Keith.<br />

University of Evansville Turtle Program Open House and Fund Raiser.<br />

Homecoming Weekend, 2004, 2005, 2006<br />

Biology representative. United Way Drive, Fall 2004.<br />

Community Service Work<br />

Babysat for parents attending English as a second language classes. Statesboro, GA 2009.<br />

Indiana Master Naturalist Program, Wesselman Woods Nature Preserve. Lecture on Reptile biology<br />

and conservation. Feb., 2007 & 2008.<br />

UE Turtle Ambassadors, Maryann Watson’s Environmental Science Class at Memorial High School.<br />

Six hatchling Malayan box turtles on loan. 2004 - 2008.<br />

Summer Turtle Classes with E-Academy.<br />

7 th -12 th graders at U.E. Two weeks, June 2005.<br />

1 st -3 rd graders at Evansville Day School. Two weeks, June 2004<br />

Turtle display at Reptile Invasion. Wesselman Woods Nature Preserve.<br />

June 14-15, 2008.<br />

June 11-12, 2005 with students Bo and Gollsheed Ouranos.<br />

June 12-13, 2004 with students Josh Yeager and Amanda Nelson.<br />

Turtle presentations at public schools.<br />

2008. Mater Dei summer program. Contact Shelly McFall.<br />

5


2006. Holy Redeemer Day Camp. Contact Becky Smither.<br />

2006. Chandler Elementary School. 5 th grade honor’s class. Contact Mrs. Strahle.<br />

2006. Evansville Day School. Middle School. Contact Mary Ann Kraft<br />

2004: 1 presentation. 3rd graders. Contact Lisa Oldham.<br />

2003: 3 presentations. Parkview Jr. High School (Lawrenceville, IL). Contact Joan Brian.<br />

2003: 3 presentations. Sharon Elementary School. 1st & 6th grade. Contact Carol Stalker.<br />

2003: 2 presentations. Evansville Day School. 1st graders. Contact Brooke Fuchs.<br />

Professional Society Memberships<br />

IUCN/SSC Tortoise and Freshwater Turtle Specialist Group<br />

Society for Integrative and Comparative Biology<br />

Turtle Survival Alliance<br />

Professional Volunteer Work<br />

Volunteer Coordinator. 25 th Annual Sea Turtle Symposium, Philadelphia, PA. 2005.<br />

Judge (student competition). 22nd Annual Sea Turtle Symposium, Miami, FL. 2002.<br />

Volunteer Chairman. 21 th Annual Sea Turtle Symposium, Philadelphia, PA. 2001.<br />

Program Co-chair. 19 th Annual Sea Turtle Symposium, South Padre Island, TX. 1999.<br />

Session Chair. 18 th and 19 th Annual Sea Turtle Symposium, 1998, 1999.<br />

Special Course Work and Skills<br />

Reading Roundtables, Georgia Southern University, Center for Teaching Excellence. 2009.<br />

Techniques in Molecular Biology. Biotechnology Institute, Penn. State University. 2002.<br />

College Teaching (graduate level course), Dept. of Education, Texas A&M University. 1996.<br />

Teacher Enhancement Workshops for Biology Lab Instructors. Dept. Biol., TAMU. 1989.<br />

Marine Biology and Oceanography Exchange Program, Duke Marine Lab, Beaufort, NC. 1988.<br />

6


Molecular Ecology (2007) 16,<br />

<strong>17</strong><strong>–</strong>18 doi: 10.1111/j.1365-294X.2006.03252.x<br />

Blackwell Publishing Ltd<br />

NEWS AND VIEWS<br />

PERSPECTIVE ARTICLE<br />

Whose turtles are they, anyway?<br />

JEANNE A. MORTIMER, *† PETER A. MEYLAN‡<br />

and MARYDELE DONNELLY§<br />

* Department of Zoology, University of Florida, Gainesville, Florida 32611, USA, † Island Conservation Society, Victoria, Mahe,<br />

Seychelles, ‡ Natural Sciences, Eckerd College, St. Petersburg, Florida 33711, USA, § Caribbean Conservation Corporation, Gainesville,<br />

Florida 32609, USA<br />

Abstract<br />

The hawksbill turtle ( Eretmochelys imbricata),<br />

listed since 1996 by the IUCN as Critically<br />

Endangered and by the Convention on <strong>International</strong> Trade in Endangered Species (CITES)<br />

as an Appendix I species, has been the subject of attention and controversy during the past<br />

10 years due to the efforts of some nations to re-open banned international trade. The most<br />

recent debate has centred on whether it is appropriate for Cuba to harvest hawksbills from<br />

shared foraging aggregations within her national waters. In this issue of Molecular Ecology,<br />

Bowen et al.<br />

have used molecular genetic data to show that such harvests are likely to have<br />

deleterious effects on the health of hawksbill populations throughout the Caribbean.<br />

Hawksbill trade involves ‘tortoiseshell’, the translucent<br />

scales covering the hawksbill plastron and carapace, which<br />

has been considered a precious material on par with ivory<br />

and rhinoceros horn since antiquity (according to legend,<br />

Cleopatra bathed in a tub made of tortoiseshell). Efforts<br />

to scientifically manage this resource were, in the past,<br />

hobbled by ignorance about key aspects of the hawksbill’s<br />

life cycle, most notably delayed sexual maturity (20<strong>–</strong><br />

30 years in the Caribbean and longer elsewhere) and the<br />

migratory nature of the species. Flipper tags, used in the<br />

context of much-reduced hawksbill populations and<br />

widely dispersed foraging habitats, revealed little about<br />

hawksbill migrations. Molecular genetics, however, have<br />

provided the technological breakthrough needed (Bass<br />

et al.<br />

1996). The study of Bowen et al.<br />

(2006) expands on that<br />

earlier work adding significantly to our understanding of<br />

the migrations of immature hawksbills. The authors use<br />

proven molecular techniques to examine the nesting beach<br />

origins of hawksbills on eight foraging grounds, four from<br />

previous literature and four newly sampled. The feeding<br />

grounds studied span the tropical West Atlantic from<br />

southern Texas to the US Virgin Islands, and include<br />

Inagua in the Bahamas and a site on the south coast of the<br />

Dominican Republic.<br />

The authors report that populations from 8 of 10 Caribbean<br />

nesting beaches studied are sufficiently distinct that<br />

their contribution to feeding aggregations can be estimated.<br />

The feeding aggregations show less genetic differentiation<br />

than nesting beaches but still exhibit significant differences.<br />

© 2006 The Authors<br />

Journal compilation © 2006 Blackwell Publishing Ltd<br />

Both maximum-likelihood and Bayesian mixed-stock analyses<br />

are used and yield similar results. Each of the feeding<br />

grounds (except Texas, which is the only sample consisting<br />

of pelagic-phase turtles) has significant contributions from<br />

multiple nesting beaches. Further analysis of the Bayesian<br />

results reveals strong relationships between feeding grounds<br />

and nesting beach populations: a direct relationship with<br />

nesting population size and an inverse relationship with<br />

distance between nesting and feeding sites. The management<br />

question that these authors ultimately address is how<br />

harvest of this species in one nation might affect populations<br />

elsewhere. Their mixed-stock analysis strongly corroborates<br />

evidence from tag returns, satellite tracking and<br />

previous genetic study, that harvests in any part of the<br />

Caribbean impact the species throughout the region.<br />

These results have important implications for conservation<br />

and for CITES. During 1970<strong>–</strong>92, Japan imported<br />

∼405<br />

<strong>17</strong>8 kg of ‘bekko’ (tortoiseshell) from 25 countries in<br />

Atlantic Latin America and the Caribbean (Milliken &<br />

Tokunaga 1987; Japanese Trade Statistics). This is equivalent<br />

to some 302 371 Caribbean hawksbill turtles (1.34 kg/<br />

turtle). By the mid-1980s many countries had acceded to<br />

CITES and stopped exporting shell, but Japan took a CITES<br />

reservation or exception on hawksbills, and continued to<br />

import shell until 1992 when international pressure forced<br />

her to drop the reservation and stop importing bekko. During<br />

those 22 years, Cuba harvested about 5000 turtles<br />

annually on their foraging grounds, contributing ∼33%<br />

of<br />

the total shell imports from the region (Carrillo et al.<br />

1999).


18 NEWS AND VIEWS<br />

By 1995, in response to the Japanese cessation of trade,<br />

Cuba reduced its annual official harvest to less than 500<br />

large turtles (Carrillo et al.<br />

1999), a level that continues to<br />

the present day. Since 1993, Cuba has stockpiled the shell<br />

from the harvest in the hope of eventually being able to sell<br />

it to Japan.<br />

In 1997 and 2000, Cuba’s interest in re-opening trade for<br />

‘Cuban hawksbills’ was presented to the biennial meeting<br />

of CITES. Cuba initially argued that hawksbills found in<br />

its waters were nonmigratory, part of a closed system and<br />

sought the rights to harvest 500 animals a year into perpetuity<br />

for the international trade and to sell off the stockpile<br />

of raw shell accumulated since 1993. Although these proposals<br />

were rejected, Cuba maintained an annual harvest<br />

of 500 or fewer large turtles. In early 2000 the stockpile was<br />

6900 kg (Cuban CITES proposal, 2000); since that time it<br />

has been increased by at ∼450<br />

turtles a year and is now<br />

estimated at ∼11<br />

200 kg of shell, representing some 7079<br />

Cuban-harvested turtles (1.59 kg/turtle).<br />

Meanwhile, Caribbean hawksbill populations are<br />

seriously depleted from their historic levels, largely as a<br />

result of centuries of trade; many populations in the region<br />

continue to decline (Mortimer & Donnelly, in press). From<br />

the 1970s onward, increasing numbers of Caribbean countries<br />

have enacted legislation that protects hawksbills on<br />

their beaches and in their waters. At some sites protective<br />

measures appear to have halted the decline of depleted populations<br />

and nesting numbers have stabilized. Remarkably,<br />

since the early 1990s, several such nesting populations<br />

increased significantly. These increases coincide with the<br />

90% decrease in the Cuban hawksbill fishery since 1994, a<br />

period during which more than 55 000 large hawksbills<br />

were spared from slaughter. This is significant given that<br />

fewer than 5000 female hawksbills are estimated to nest<br />

annually in the Caribbean region (Meylan & Donnelly<br />

1999; Mortimer & Donnelly, in press). The implications of<br />

the paper by Bowen et al.<br />

are that harvest of hawksbills<br />

from any nesting beach could impact multiple foreign<br />

feeding grounds, and that harvest on any feeding grounds<br />

could impact the nesting beach populations at multiple<br />

sites. Thus, any nation that opens or promotes harvest<br />

could be undermining badly needed efforts to conserve the<br />

species at other sites.<br />

There is more at stake than turtles. Hawksbills are spongivores<br />

(Leon & Bjorndal 2002) and spongivory helps to<br />

maintain healthy coral reefs by releasing corals from competition<br />

with sponges (Bjorndal & Jackson 2003). Historic<br />

consumption of sponges by hawksbill turtles is estimated<br />

to have been 800 times higher than it is now (McClenachan<br />

et al.<br />

2006). At a time when coral reefs are among the most<br />

endangered ecosystems on the planet (Wilkinson 2000),<br />

successful conservation and management of hawksbills<br />

may be an essential component for ecosystem restoration.<br />

References<br />

Bass AL, Good DA, Bjorndal KA et al. (1996) Testing models of<br />

female reproductive migratory behavior and population structure<br />

in the Caribbean hawksbill turtle, Eretmochelys imbricata,<br />

with mtDNA sequences. Molecular Ecology,<br />

5,<br />

321<strong>–</strong>328.<br />

Bjorndal KA, Jackson JBC (2003) Role of sea turtles in marine ecosystems<br />

— reconstructing the past. In: Biology of Sea Turtles (eds<br />

Lutz PL, Musick JA, Wyneken J), Vol. II, pp. 259<strong>–</strong>273. CRC Press,<br />

Boca Raton, Florida.<br />

Bowen BW, Grant WS, Hillis-Starr Z et al. (2006) Mixed-stock<br />

analysis reveals the migrations of juvenile hawksbill turtles<br />

( Eretmochelys imbricata)<br />

in the Caribbean Sea. Molecular Ecology,<br />

doi:10.1111/j.1365-294X.2006.03096.x.<br />

Carrillo CE, Webb GJW, Manolis SC (1999) Hawksbill turtles<br />

( Eretmochelys imbricata)<br />

in Cuba: an assessment of the historical<br />

harvest and its impacts. Chelonian Conservation and Biology,<br />

3,<br />

264<strong>–</strong>280.<br />

Leon YM, Bjorndal KA (2002) Selective feeding in the hawksbill<br />

turtle, an important predator in coral reef ecosystems. Marine<br />

Ecology Progress Series,<br />

245,<br />

249<strong>–</strong>258.<br />

McClenachan L, Jackson JBC, Newman MJH (2006) Conservation<br />

implications of historic sea turtle nesting beach loss. Front Ecological<br />

Environment,<br />

4,<br />

290<strong>–</strong>296.<br />

Meylan AB, Donnelly M (1999) Status justification for listing the<br />

hawksbill turtle ( Eretmochelys imbricata)<br />

as critically endangered<br />

on the 1996 IUCN Red List of Threatened Animals. Chelonian<br />

Conservation and Biology,<br />

3,<br />

<strong>17</strong>7<strong>–</strong>184.<br />

Milliken T, Tokunaga H (1987) The Japanese sea turtle trade<br />

1970<strong>–</strong>86. A Special Report Prepared by TRAFFIC (Japan) . Center<br />

for Environmental Education, Washington D.C.<br />

Mortimer JA, Donnelly M (in press)). IUCN 2006 Global Assessment<br />

for the Hawksbill Turtle (Eretmochelys imbricata) .<br />

Wilkinson CR (2000) Status of Coral Reefs of the World: 2000.<br />

Global<br />

Coral Reef Monitoring Network.<br />

Australian Institute of Marine<br />

Science.<br />

© 2006 The Authors<br />

Journal compilation © 2006 Blackwell Publishing Ltd


Molecular Ecology (2009) 18, 4841<strong>–</strong>4853 doi: 10.1111/j.1365-294X.2009.04403.x<br />

Turtle groups or turtle soup: dispersal patterns of<br />

hawksbill turtles in the Caribbean<br />

J. M. BLUMENTHAL,*† F. A. ABREU-GROBOIS,‡ T. J. AUSTIN,* A. C. BRODERICK,†<br />

M. W. BRUFORD,§ M. S. COYNE,†<strong>–</strong> G. EBANKS-PETRIE,* A. FORMIA,§ P. A. MEYLAN,**<br />

A. B. MEYLAN†† and B. J. GODLEY†<br />

*Department of Environment, Box 486, Grand Cayman KY1-1106, Cayman Islands, †Centre for Ecology and Conservation,<br />

School of Biosciences, University of Exeter Cornwall Campus, Penryn TR10 9EZ, UK, ‡Laboratorio de Genética, Unidad<br />

Académica Mazatlán, Instituto de Ciencias del Mar y Limnología, Mazatlán, Sinaloa 82040, México, §Biodiversity and<br />

Ecological Processes Research Group, School of Biosciences, Cardiff University, Cardiff CF10 3TL, UK, <strong>–</strong>SEATURTLE.ORG,<br />

Durham, NC 27705, USA, **Natural Sciences, Eckerd College, St Petersburg, FL 33711, USA, ††Florida Fish and Wildlife<br />

Conservation Commission, Fish and Wildlife Research Institute, St Petersburg, FL 33701, USA<br />

Introduction<br />

Abstract<br />

Despite intense interest in conservation of marine turtles, spatial ecology during the<br />

oceanic juvenile phase remains relatively unknown. Here, we used mixed stock analysis<br />

and examination of oceanic drift to elucidate movements of hawksbill turtles (Eretmochelys<br />

imbricata) and address management implications within the Caribbean. Among<br />

samples collected from 92 neritic juvenile hawksbills in the Cayman Islands we detected<br />

11 mtDNA control region haplotypes. To estimate contributions to the aggregation, we<br />

performed ‘many-to-many’ mixed stock analysis, incorporating published hawksbill<br />

genetic and population data. The Cayman Islands aggregation represents a diverse mixed<br />

stock: potentially contributing source rookeries spanned the Caribbean basin, delineating<br />

a scale of recruitment of 200<strong>–</strong>2500 km. As hawksbills undergo an extended phase of<br />

oceanic dispersal, ocean currents may drive patterns of genetic diversity observed on<br />

foraging aggregations. Therefore, using high-resolution Aviso ocean current data, we<br />

modelled movement of particles representing passively drifting oceanic juvenile<br />

hawksbills. Putative distribution patterns varied markedly by origin: particles from<br />

many rookeries were broadly distributed across the region, while others would appear to<br />

become entrained in local gyres. Overall, we detected a significant correlation between<br />

genetic profiles of foraging aggregations and patterns of particle distribution produced<br />

by a hatchling drift model (Mantel test, r = 0.77, P < 0.001; linear regression, r = 0.83,<br />

P < 0.001). Our results indicate that although there is a high degree of mixing across the<br />

Caribbean (a ‘turtle soup’), current patterns play a substantial role in determining genetic<br />

structure of foraging aggregations (forming turtle groups). Thus, for marine turtles and<br />

other widely distributed marine species, integration of genetic and oceanographic data<br />

may enhance understanding of population connectivity and management requirements.<br />

Keywords: conservation genetics, hawksbill, marine turtle, migratory species, mixed stock,<br />

ocean currents<br />

Received 29 June 2009; revision received 21 September 2009; accepted 25 September 2009<br />

Many marine vertebrates have life cycles that span wide<br />

spatiotemporal scales <strong>–</strong> complicating research and<br />

Correspondence: Brendan J. Godley, Fax: 44 (0) 1326 253638;<br />

E-mail: b.j.godley@exeter.ac.uk<br />

Ó 2009 Blackwell Publishing Ltd<br />

management. Recently, molecular techniques have provided<br />

insights into patterns of migration and stock resolution<br />

in species ranging from fish (Millar 1987) to<br />

porpoises (Andersen et al. 2001) and great whales<br />

(Baker et al. 1999; Witteveen et al. 2004). Such stock resolution<br />

issues are particularly important when species


4842 J. M. BLUMENTHAL ET AL.<br />

are commercially valuable, cross geopolitical boundaries<br />

or have been exploited to the point of endangerment.<br />

Marine turtles exemplify these concerns: a long history<br />

of commercial use has resulted in global declines<br />

(Baillie et al. 2004) and management of marine turtle<br />

populations is complex, as developmental and reproductive<br />

migrations may span the territorial waters of<br />

several nations and the high seas (Bowen et al. 1995;<br />

Bolten et al. 1998).<br />

Like many other marine organisms (Mora & Sale<br />

2002) most marine turtle species experience a phase of<br />

oceanic dispersal, which in marine turtles is known as<br />

the ‘lost year’ or ‘lost years’ (Carr 1987). This extended<br />

period of drifting in concert with initially small size<br />

and high mortality inhibits tracking movements of neonates<br />

from the time of entry into marine habitat off<br />

natal beaches to the time of recruitment into neritic foraging<br />

grounds (Bolten et al. 1998; Lahanas et al. 1998;<br />

Engstrom et al. 2002; Bowen et al. 2004; Luke et al.<br />

2004). Although foraging grounds are typically inhabited<br />

by individuals originating from multiple nesting<br />

beaches a behavioural barrier to genetic mixing at nesting<br />

beaches is maintained: generations of mature<br />

females return to their natal regions to reproduce (Meylan<br />

et al. 1990). This ‘natal homing’ leads to genetic differentiation<br />

at mitochondrial loci of nesting populations<br />

over time (hawksbills Eretmochelys imbricata: Broderick<br />

et al. 1994; Bass et al. 1996; loggerheads Caretta caretta:<br />

Encalada et al. 1998; Engstrom et al. 2002; green turtles<br />

Chelonia mydas: Meylan et al. 1990; Allard et al. 1994).<br />

Therefore, mitochondrial DNA haplotypes and haplotype<br />

frequencies characteristic of each region serve as<br />

a form of ‘genetic tag,’ linking juveniles in mixed<br />

aggregations with their specific nesting beach origins<br />

(Norman et al. 1994).<br />

In a method originally developed to resolve the origins<br />

of salmon stocks (Millar 1987), estimates of marine<br />

turtle origin can be made using a maximum likelihood<br />

(Pella & Milner 1987) or Bayesian (Pella & Masuda<br />

2001) mixed stock analysis (MSA), where haplotype frequencies<br />

for genetically mixed aggregations are compared<br />

with potential source populations. MSA has<br />

recently proven valuable in elucidating migratory patterns<br />

and range states in hawksbill (Bowen et al. 1996,<br />

2007a; Bass 1999; Diaz-Fernandez et al. 1999; Troëng<br />

et al. 2005; Velez-Zuazo et al. 2008), loggerhead (Bowen<br />

et al. 1995; Laurent et al. 1998; Engstrom et al. 2002;<br />

Maffucci et al. 2006) and green turtles (Lahanas et al.<br />

1998; Luke et al. 2004). While mixed stock analysis has<br />

traditionally been used to estimate contributions of<br />

many potential source rookeries to a single foraging<br />

ground (‘many-to-one’ analysis), a new approach has<br />

recently been developed which more realistically estimates<br />

the contributions of many rookeries to many for-<br />

aging grounds within a metapopulation context (‘manyto-many’<br />

analysis; Bolker et al. 2007). Caribbean hawksbills<br />

represent ideal candidates for many-to-many<br />

analysis: rookeries are sufficiently differentiated (Bass<br />

et al. 1996), populations may be relatively contained in<br />

the Caribbean region (Bowen et al. 1996) and data from<br />

multiple mixed stocks have recently been published<br />

(Bowen et al. 2007a; Velez-Zuazo et al. 2008).<br />

Tagging and satellite tracking have demonstrated<br />

that hawksbill developmental and reproductive migrations<br />

span the territorial waters of multiple jurisdictions<br />

(Meylan 1999; Horrocks et al. 2001; Troëng et al.<br />

2005; Whiting & Koch 2006; van Dam et al. 2008) and<br />

genetic research is beginning to elucidate links<br />

between nesting beaches and foraging grounds (Bowen<br />

et al. 1996, 2007a; Bass 1999; Diaz-Fernandez et al.<br />

1999; Troëng et al. 2005; Velez-Zuazo et al. 2008). It<br />

has also been determined that Caribbean hawksbill<br />

foraging aggregations show shallow but significant<br />

genetic structure (Bowen et al. 2007a). As in green turtles<br />

(Bass et al. 2006), this suggests that rather than the<br />

oceanic phase being a homogeneous mixture (the ‘turtle<br />

soup’ model: Engstrom et al. 2002), dispersal is<br />

non-random. However for many populations links<br />

between nesting beaches and foraging areas have not<br />

been identified (Bowen et al. 1996) and little is known<br />

regarding movements of oceanic juveniles during the<br />

lost year or years (Musick & Limpus 1997; Bolten<br />

2003) and factors which drive dispersal during this<br />

phase (Bowen et al. 2007a).<br />

In previous genetic studies of green (Lahanas et al.<br />

1998; Bass & Witzell 2000; Luke et al. 2004) and loggerhead<br />

turtles (Engstrom et al. 2002; Reece et al. 2006)<br />

source population size and geographic distance have<br />

proven inconsistent in determining recruitment from<br />

rookeries to foraging grounds. For Caribbean hawksbills,<br />

these factors were significantly correlated with foraging<br />

aggregation genetic composition, but correlations<br />

were weak and significant only for a small proportion<br />

of individual foraging grounds (Bowen et al. 2007a).<br />

Ocean currents have been postulated as playing a major<br />

role in distribution of oceanic juvenile marine turtles<br />

(Carr 1987) and recent studies have linked genetic composition<br />

of marine turtle foraging aggregations to ocean<br />

current patterns (Luke et al. 2004; Okuyama & Bolker<br />

2005; Bass et al. 2006; Carreras et al. 2006). However,<br />

for hawksbills, attempts to account for the role of ocean<br />

currents through incorporation of a current correction<br />

factor (increasing or decreasing geographic distances in<br />

a regression model by 10<strong>–</strong>20%) did not increase the<br />

strength of correlations (Bowen et al. 2007a).<br />

To evaluate population connectivity for marine organisms,<br />

empirical data must be linked with biophysical<br />

modelling of oceanic dispersal (Botsford et al. 2009).<br />

Ó 2009 Blackwell Publishing Ltd


Generalized ocean current diagrams can aid in consideration<br />

of the role of ocean currents in marine turtle<br />

dispersal (Bass et al. 2006; Carreras et al. 2006; Bowen<br />

et al. 2007a) but this method falls short when current<br />

patterns are complex. Trans-Atlantic drift has been<br />

modelled for oceanic juvenile loggerheads (Hays &<br />

Marsh 1997) and the availability of high-resolution<br />

ocean current data for the Caribbean and other regions<br />

now opens up the possibility of modelling dispersal of<br />

oceanic juvenile hawksbills in dynamic ocean currents.<br />

In this study, characterization of stock composition at<br />

the Cayman Islands hawksbill foraging site was<br />

extended by development of a hatchling drift model<br />

incorporating source population sizes and regional<br />

ocean current data to answer fundamental questions of<br />

hawksbill ecology: (i) how are oceanic juvenile hawksbills<br />

distributed during the lost year or years? (ii) how<br />

are foraging stocks geographically structured or regionally<br />

mixed? (iii) what role might ocean currents play in<br />

determining these patterns?<br />

Materials and Methods<br />

Study area<br />

The Cayman Islands are located in the Caribbean Sea,<br />

approximately 240 km south of Cuba (Fig. 1). The<br />

three low-lying islands are exposed peaks on the<br />

Cayman Ridge formation, with nearly vertical slopes<br />

extending to depths in excess of 2000 m on all sides<br />

(Roberts 1994). Hawksbill nesting, while described as<br />

abundant in historical records (Lewis 1940), appears to<br />

have been largely extirpated in recent decades (Bell<br />

et al. 2007). However, the islands host foraging aggregations<br />

of neritic juvenile hawksbill turtles, inhabiting<br />

colonized pavement, coral reef and reef wall habitats<br />

(Blumenthal et al. 2009a, b). For this study, two sampling<br />

sites were selected: Bloody Bay, Little Cayman<br />

(19.7°N, 81.1°W) and western Grand Cayman (19.3°N,<br />

81.4°W).<br />

Capture and sampling<br />

Juvenile hawksbills (20<strong>–</strong>60 cm straight carapace length)<br />

were hand-captured in neritic foraging habitat and flipper<br />

and PIT tagged according to standard protocols<br />

(Blumenthal et al. 2009b) to prevent repeated sampling<br />

of individuals. Genetic sampling for this study was<br />

undertaken year-round between 2000 and 2003. Skin<br />

biopsies were obtained from a rear flipper with a sterile<br />

4 mm biopsy punch and preserved in a solution of 20%<br />

DMSO saturated with NaCl (Dutton 1996). Blood samples<br />

were collected from the dorsal cervical sinus<br />

(Owens & Ruiz 1980) and preserved in lysis buffer<br />

Ó 2009 Blackwell Publishing Ltd<br />

HAWKSBILL TURTLE DISPERSAL 4843<br />

Fig. 1 Distribution of hawksbill aggregations studied in the<br />

Caribbean region. Circles represent locations of Caribbean<br />

hawksbill rookeries: filled circles are genetically characterized.<br />

Haplotype frequency data from Antigua, Barbados, Belize,<br />

Costa Rica, Cuba, Mexico, Puerto Rico and the US Virgin<br />

Islands were used in many-to-many mixed stock analysis;<br />

rookeries in Venezuela are presently insufficiently characterized<br />

to permit inclusion. Triangles represent genetically characterized<br />

hawksbill foraging grounds also incorporated in mixed<br />

stock analysis: Bahamas, Cayman, Cuba, Dominican Republic,<br />

Mexico, Puerto Rico, Texas and the US Virgin Islands. Marine<br />

ecoregions (1) Gulf of Mexico (2) western Caribbean (3) southwestern<br />

Caribbean (4) Greater Antilles (5) southern Caribbean<br />

(6) eastern Caribbean (7) Bahamian. Source of published<br />

genetic data: Antigua (Bass 1999), Barbados (Browne et al. in<br />

press), Belize (Bass 1999), Cuba (Diaz-Fernandez et al. 1999),<br />

Mexico (Bass 1999; Diaz-Fernandez et al. 1999), USVI (Bass<br />

1999; Bowen et al. 2007a), Costa Rica (Troëng et al. 2005;<br />

Bowen et al. 2007a) and Puerto Rico (Diaz-Fernandez et al.<br />

1999; Velez-Zuazo et al. 2008).<br />

(100mM Tris-HCl, pH 8, 100mM EDTA, pH 8, 10mM<br />

NaCl and 1<strong>–</strong>2% SDS: Dutton 1996).<br />

Molecular analysis<br />

Following overnight digestion at 55 °C with proteinase<br />

K, DNA extraction was conducted via standard phenol<br />

chloroform extraction (Milligan 1998), Qiagen DNEasy<br />

tissue kit, or a modification of a protocol by Allen et al.<br />

(1998) (Formia et al. 2006, 2007). Fragments of 486 or<br />

550 base pairs (bp) were amplified via polymerase<br />

chain reaction (PCR), using primer pairs TCR6 (Norman<br />

et al. 1994) ⁄ L15926 (Kocher et al. 1989) or LTEi3 ⁄ HDEi1<br />

(LTEi3 5¢-CCTAGAATAATCAAAAGAGAAGG-3¢;<br />

HDEi1 5¢- AGTTTCGTTAATTCGGCAG-3¢; Abreu-Grobois<br />

et al. 2006) respectively. PCR reactions [PCR<br />

Ready-To-Go Beads (Amersham) or 1.5 mM MgCl2, 1·<br />

PCR buffer, 200 lM of each dNTP, 0.5 lM of each<br />

primer, 0.5 U Invitrogen Taq DNA polymerase, 1 lL


4844 J. M. BLUMENTHAL ET AL.<br />

template DNA and sterile H2O to a volume of 10 lL)<br />

were carried out in a Perkin Elmer GeneAmp PCR<br />

system 9700 (3 min at 94 °C, 35 cycles of: 45 s at<br />

94 °C, 30 s at 55 °C and 1.5 min at 72 °C, followed by<br />

72 °C for 10 min; Formia et al. 2006). PCR products<br />

were cleaned with a Geneclean Turbo Kit (Qbiogene)<br />

or QIAquick spin columns (Qiagen), sequenced in<br />

both directions with a BigDye Terminator Cycle<br />

Sequencing Kit v. 2.0 (Applied Biosystems), purified<br />

via ethanol precipitation and run on an Applied Biosystems<br />

model 3100 automated DNA sequencer at<br />

Cardiff University, or analysed at the University of<br />

Florida Sequencing Core on an Amersham Pharmacia<br />

Biotech MegaBACE 1000 capillary array DNA<br />

sequencing unit. Negative controls (template-free PCR<br />

reactions) were utilized throughout to test for contamination.<br />

To identify haplotypes, sequences were aligned with<br />

published hawksbill mtDNA sequences (Bass et al.<br />

1996; Bowen et al. 1996; Bass 1999; Diaz-Fernandez<br />

et al. 1999; Troëng et al. 2005; Bowen et al. 2007a;<br />

Velez-Zuazo et al. 2008; Browne et. al. in press) in Sequencher<br />

2.1.2 (Gene Codes Corp). Haplotype diversity<br />

(h), nucleotide diversity (p, and population pairwise<br />

FST values (10 000 permutations; Kimura 2-parameter<br />

model, a = 0.50) were calculated in Arlequin v. 3.11<br />

(Excoffier et al. 2005).<br />

Mixed stock analysis<br />

Weighted and unweighted many-to-many analyses (Bolker<br />

et al. 2007) were also conducted, incorporating all<br />

previously published Caribbean hawksbill genetic data<br />

(haplotype frequencies: Antigua (Bass 1999), Barbados<br />

(Browne et al. in press; windward and leeward rookeries<br />

combined), Belize (Bass 1999), Cuba (Diaz-Fernandez<br />

et al. 1999; all foraging grounds combined), Mexico<br />

(Bass 1999; Diaz-Fernandez et al. 1999), USVI (Bass<br />

1999; Bowen et al. 2007a), Costa Rica (Troëng et al.<br />

2005; Bowen et al. 2007a) and Puerto Rico (Diaz-Fernandez<br />

et al. 1999; Velez-Zuazo et al. 2008). To enable comparison<br />

with previously published rookery haplotypes,<br />

sequences were truncated to 384 bp for analysis. Source<br />

population size (number of nests) was used as a strict<br />

constraint in the ‘weighted’ many-to-many model<br />

(Bolker et al. 2007), with population sizes taken from<br />

Mortimer & Donnelly (2007).<br />

Hatchling drift model<br />

Movement of particles was modelled using high resolution<br />

ocean current data from Aviso (Archiving, Validation<br />

and Interpretation of Satellite Oceanographic data,<br />

http://www.aviso.oceanobs.com). Aviso geostrophic<br />

velocity vector (GVV) data (maps of absolute geostrophic<br />

velocities derived from maps of Absolute<br />

Dynamic Topography combining all satellites, obtained<br />

from http://www.aviso.oceanobs.com) have a spatial<br />

resolution of 1 ⁄ 3 degree and a temporal resolution of<br />

1 day. To model the movement of passively drifting<br />

hawksbill hatchlings, post-hatchlings and oceanic juveniles,<br />

virtual particles were released from the coordinates<br />

of genetically characterized Caribbean hawksbill<br />

rookeries locations (Fig. 1). To simulate peak hatchling<br />

emergence, particles were released 60 days after the<br />

peak of the nesting season at each location (nesting season<br />

data: Chacón 2004; Garduño-Andrade 1999; Moncada<br />

et al. 1999), with one particle released per day over a<br />

60-day period and the resulting particle profiles<br />

weighted by source population size (taken from Mortimer<br />

& Donnelly 2007). Drift of virtual particles was initiated<br />

approximately 50 km offshore from each rookery to<br />

account for the phase of directed hatchling swimming<br />

that occurs prior to beginning passive drift (Hasbún<br />

2002; Chung et al. 2009a,b) and to bring particles into<br />

the area covered by Aviso data, as nearshore currents<br />

are too complex to map reliably. The u (eastward) and v<br />

(northward) ocean current vector components were sampled<br />

at the location of each particle from the daily GVV<br />

data file using a bilinear interpolation [Generic Mapping<br />

Tools (GMT 4.11; Wessel & Smith 1991)]. The u and v<br />

values (cm ⁄ s) were converted to m ⁄ day and the particle<br />

advected that distance in the x and y direction, respectively,<br />

and a new location obtained for that particle. Particles<br />

that left the GVV field (i.e. pushed towards the<br />

coastline) were removed from the model, as there is no<br />

information on how an oceanic juvenile hawksbill would<br />

behave in this situation. Also, while natural mortality<br />

during the hatchling, post-hatchling and oceanic juvenile<br />

phase is high, this factor was not incorporated into the<br />

model. One location per day was saved for each particle<br />

for analysis. All current modelling was carried out using<br />

custom perl scripts, the geod program, part of the<br />

PROJ.4 Cartographic Projections Library (http://<br />

www.remotesensing.org/proj/) and the GMT package.<br />

Comparison of the hatchling drift model with mixed<br />

stock analysis<br />

To examine the possible influence of ocean currents on<br />

patterns of dispersal, we compared foraging aggregation<br />

genetic profiles (determined via many-to-many<br />

analysis) with patterns of particle distribution (particle<br />

profiles) produced by the hatchling drift model<br />

(Table 1). To produce particle profiles, we used the Global<br />

Maritime Boundaries Database (GMBD) to delineate<br />

Exclusive Economic Zones of nations containing genetically<br />

characterized hawksbill foraging aggregations.<br />

Ó 2009 Blackwell Publishing Ltd


Table 1 Comparison of genetic profiles from many-to-many analysis and particle profiles derived from the hatchling drift model<br />

We then quantified the number of particles from each<br />

source to enter these territorial waters during the modelled<br />

first year of drifting (though it should be noted that<br />

a longer oceanic phase could occur; Musick & Limpus<br />

1997). Euclidean distance matrices were calculated for<br />

particle and genetic profiles (treating each estimate of<br />

proportional contribution as a character) and compared<br />

with a Mantel test (PopTools v. 3.0; Hood 2009). Additionally,<br />

to facilitate comparison with the results of<br />

Bowen et al. (2007a), we investigated the relationship<br />

between genetic and particle profiles with linear regression,<br />

regressing arcsine transformed proportions (Graph-<br />

Pad InStat 3.10, GraphPad Software).<br />

Results<br />

Cay Tex Bah Cuba PR USVI DR Mex<br />

MSA Drift MSA Drift MSA Drift MSA Drift MSA Drift MSA Drift MSA Drift MSA Drift<br />

Anti 6.0 13.1 1.9 0.0 5.2 13.2 2.7 4.3 5.3 18.5 10.2 50.2 4.4 23.1 6.7 3.8<br />

Barb 46.1 40.3 3.3 0.0 22.7 25.9 15.7 13.5 47.4 28.6 27.7 21.2 65.7 35.3 5.8 13.8<br />

Belz 0.7 0.0 0.6 0.0 0.8 0.0 0.9 0.0 1.5 0.0 0.7 0.0 2.1 0.0 0.7 0.6<br />

Cuba 21.9 45.7 2.4 0.0 26.1 0.0 44.3 64.6 8.0 0.0 23.0 0.0 10.1 0.0 5.8 0.0<br />

Mexi 13.2 0.0 85.3 100.0 28.4 46.6 10.5 <strong>17</strong>.1 8.5 0.0 14.1 0.0 7.7 0.0 71.6 81.3<br />

USVI 3.0 0.3 2.0 0.0 3.7 0.7 2.4 0.2 8.1 8.5 2.3 26.4 3.9 4.0 2.5 0.2<br />

CR 0.5 0.6 0.3 0.0 0.5 0.5 0.7 0.3 1.4 0.0 0.4 0.0 1.5 0.0 0.4 0.3<br />

PR 8.7 0.0 4.1 0.0 12.6 13.0 22.9 0.0 19.7 44.4 21.6 2.1 4.7 37.6 6.3 0.0<br />

Rows represent potential source rookeries <strong>–</strong> Antigua, Barbados, Belize, Cuba, Mexico, US Virgin Islands, Costa Rica and Puerto Rico;<br />

columns represent genetically characterized foraging grounds <strong>–</strong> Cayman Islands, Texas, Bahamas, Cuba, Puerto Rico, US Virgin<br />

Islands, Dominican Republic and Mexico (for sources of published genetic data see Fig. 1). Numbers are mean proportional<br />

contribution from each rookery to each foraging ground, estimated via weighted many-to-many mixed stock analysis (MSA) and<br />

proportion of particles from each rookery to enter each foraging ground during the first year of drifting, from the hatchling drift<br />

model (Drift).<br />

Genetic structure<br />

Among 92 neritic juvenile hawksbills from the Cayman<br />

Islands, we detected 11 mitochondrial DNA (mtDNA)<br />

control region haplotypes: Ei-A1, Ei-A2, Ei-A3, Ei-A9c,<br />

Ei-A11c, Ei-A18, Ei-A20, Ei-A24, EiA28, Ei-A29 and a<br />

previously undetected haplotype, Ei-A72 (GenBank<br />

accession number GQ925509). The Cayman Islands foraging<br />

aggregation represented a diverse mixed stock<br />

(h = 0.72 ± 0.04; p = 0.01 ± 0.005). Haplotype frequencies<br />

from Grand Cayman and Little Cayman were not<br />

significantly different (population pairwise FST = )0.01,<br />

P = 0.83). Therefore results for the two study sites were<br />

pooled for subsequent analyses.<br />

Rookeries of origin<br />

Using unweighted and weighted many-to-many analysis,<br />

we estimated the contributions of genetically character-<br />

Ó 2009 Blackwell Publishing Ltd<br />

HAWKSBILL TURTLE DISPERSAL 4845<br />

ized Caribbean rookeries to the Cayman Islands foraging<br />

aggregation. Shrink factors for all chains were


4846 J. M. BLUMENTHAL ET AL.<br />

Fig. 2 Genetic composition at the Cayman Islands foraging<br />

aggregation. Among the 92 individuals sequenced, 11 haplotypes<br />

were detected: Ei-A1 (44); Ei-A2 (2); Ei-A3 (1); Ei-A9c (8);<br />

Ei-A11c (<strong>17</strong>); Ei-A18 (1); Ei-A20 (1); Ei-A24 (11); Ei-A28 (4); Ei-<br />

A29 (1); Ei-A72 (2). (a) Using an unweighted many-to-many<br />

model (white bars), estimated contributions (mean proportion,<br />

2.5% and 97.5% confidence intervals) of genetically characterized<br />

hawksbills rookeries to the Cayman Islands foraging aggregation<br />

were Antigua (21.3%, 3.0<strong>–</strong>41.8%), Barbados (20.5%,<br />

1.1<strong>–</strong>48.1%), Belize (8.7%, 0.4<strong>–</strong>24.9%), Cuba (12.6%, 0.6<strong>–</strong>34.1%),<br />

Mexico (9.1%, 0.8<strong>–</strong>16.0%), USVI (11%, 0.4<strong>–</strong>27.6%), Costa Rica<br />

(10.7%, 2.9<strong>–</strong>21.3%), Puerto Rico (6.1%, 0.8<strong>–</strong>16.6%). When<br />

source population size was used as a constraint (grey bars), estimates<br />

were Antigua (6.0%, 0.7<strong>–</strong>15.5%), Barbados (46.1%, 20.6<strong>–</strong><br />

69.0%), Belize (0.7%, 0.0<strong>–</strong>2.7%), Cuba (21.9%, 3.7<strong>–</strong>43%), Mexico<br />

(13.2%, 7.4<strong>–</strong>20.6%), USVI (3%, 0.1<strong>–</strong>9.4%), Costa Rica (0.5%,<br />

0.0<strong>–</strong>2.0%), Puerto Rico (8.7%, 1.5<strong>–</strong>18.3%). (b) Arrows are scaled<br />

in proportion to estimated mean contribution (weighted manyto-many)<br />

to indicate potential scale of recruitment.<br />

365 days of drifting) and genetic profiles (many-tomany<br />

results for all genetically characterized Caribbean<br />

foraging aggregations) (Fig. 4).<br />

Discussion<br />

Because hawksbill turtles spend the vast majority of<br />

their lives at sea, for many populations links between<br />

nesting beaches and foraging grounds are unknown.<br />

To estimate origins of neritic juvenile hawksbill turtles<br />

foraging in the Cayman Islands, we applied many-tomany<br />

modelling using data from this study and other<br />

published genetic studies of Caribbean hawksbills. For<br />

the Cayman Islands foraging aggregation potentially<br />

contributing source rookeries spanned the Caribbean<br />

basin <strong>–</strong> delineating a scale of recruitment of 200<strong>–</strong><br />

2500 km (straight-line distance) and highlighting the<br />

complex spatial ecology of the species.<br />

We then investigated the role of ocean currents as a<br />

mechanism underlying patterns of dispersal. Bowen<br />

et al. (2007a) previously tested the role of geographic<br />

distance (both straight-line and modified in order to<br />

account for a likely influence of ocean currents) in determining<br />

recruitment of hawksbills from nesting populations<br />

to foraging aggregations. Here, we compared<br />

results of mixed stock analysis to particle profiles, which<br />

take source population size, geographic distance and<br />

oceanographic circulation patterns into account <strong>–</strong> an<br />

important consideration as oceanic juveniles are unlikely<br />

to follow a direct route between nesting beaches and foraging<br />

grounds. A strong correlation was detected<br />

between patterns of stock composition at Caribbean<br />

hawksbill foraging aggregations and patterns of particle<br />

distribution at these sites <strong>–</strong> indicating that ocean currents<br />

likely play a substantial role in determining geographic<br />

patterns of genetic diversity, and that models of<br />

oceanic drift may illustrate movements of oceanic juvenile<br />

turtles during the lost year or years. Putative dispersal<br />

patterns for oceanic juvenile hawksbills varied<br />

regionally: particles from many rookeries were broadly<br />

distributed across the Caribbean (a ‘turtle soup’), while<br />

particles from other rookeries appeared to be regionally<br />

constrained (forming ‘turtle groups’). Thus, because of<br />

the complexities of regional and local current patterns,<br />

some areas will be more diverse (turtles recruiting from<br />

multiple jurisdictions) while others will experience<br />

greater levels of proximate or local recruitment.<br />

Ultimately, population resolution may allow demographic<br />

parameters such as abundance, survival, sex<br />

ratio and growth to be linked across broad spatiotemporal<br />

scales and incorporated into population modelling<br />

(Bowen et al. 2004; Reece et al. 2006). However, at present,<br />

the accuracy of mixed stock estimates is limited by<br />

insufficient baseline data from nesting beaches (Troëng<br />

et al. 2005). Indeed, detection of a previously unknown<br />

haplotype at the Cayman Islands foraging aggregation<br />

indicates that nesting beaches in the Caribbean or source<br />

rookeries farther afield have been incompletely or insufficiently<br />

sampled. Longer periods of sampling and larger<br />

sample sizes (Reece et al. 2006), standardized use of<br />

primers which amplify larger mtDNA fragments (Abreu-<br />

Grobois et al. 2006), and expanded geographic coverage<br />

Ó 2009 Blackwell Publishing Ltd


HAWKSBILL TURTLE DISPERSAL 4847<br />

Fig. 3 Trajectories of modelled particles representing drifting oceanic stage juvenile turtles, released from hawksbill index beaches<br />

in (a) Gulf of Mexico (Isla Holbox, Punta Xen and Las Coloradas Mexico) (b) Greater Antilles (Doce Leguas Cuba, Mona Island<br />

Puerto Rico) (c) eastern Caribbean (Long Beach Antigua, Hilton Beach Barbados, Buck Island USVI) (d) southern Caribbean (Archipiélago<br />

Los Roques and Península de Paria Venezuela) (e) southwestern Caribbean (Tortuguero Costa Rica) (f) western Caribbean<br />

(Gales Point Belize) (g) depicts overall particle trajectory patterns in the wider Caribbean.<br />

Ó 2009 Blackwell Publishing Ltd


4848 J. M. BLUMENTHAL ET AL.<br />

Fig. 4 Comparison of hatchling drift (black bars) with weighted (grey bars) and unweighted (white bars) many-to-many results for<br />

the Cayman Islands and other genetically characterized foraging grounds (for sources of published genetic data see Fig. 1). Data are<br />

mean proportion (±2.5<strong>–</strong>97.5% confidence intervals for many-to-many analysis).<br />

(Engstrom et al. 2002) are priorities for improving estimates.<br />

Ocean current modelling may serve to identify<br />

potential locations of regional foraging aggregations<br />

and priorities for studies of unsurveyed source rookeries:<br />

for example, particles from the southern Caribbean<br />

(Venezuela) would appear to be widely dispersed, yet<br />

rookeries are presently insufficiently characterized to<br />

permit inclusion in mixed stock estimates.<br />

Within the context of developing a hatchling drift<br />

model, much remains to be resolved or refined regarding<br />

biological parameters (e.g. sizes of hawksbill source<br />

populations, timing of peak hatching, duration of the<br />

Ó 2009 Blackwell Publishing Ltd


oceanic phase, behaviour of oceanic juveniles pushed<br />

toward landmasses, and roles of active swimming and<br />

orientation in hatchlings, post-hatchlings, and oceanic<br />

juveniles) and modelling ocean currents (e.g. incorporating<br />

wind-induced components; Hays & Marsh 1997;<br />

Girard et al. 2009). In this study, our hatchling drift<br />

model was limited to determining exposure of foraging<br />

sites to particles during the first year of drift. This is<br />

useful as a measure of potential for recruitment from<br />

each of the sources, as the timeframes within which<br />

hawksbills begin actively swimming and recruit to foraging<br />

grounds is unknown. However, a longer oceanic<br />

phase may occur (Musick & Limpus 1997), calling for<br />

extended modelling of oceanic drift when such longterm<br />

high-resolution oceanographic data become available.<br />

Comparison of multiple years of Aviso data may<br />

also be informative in determining inter-annual variations<br />

in recruitment and implications for management:<br />

if, as in green turtles, there is temporal structuring in<br />

the genetic composition of neritic juvenile foraging<br />

aggregations, then marine turtle management plans<br />

should not be based on the ‘snapshot’ of short-term<br />

genetic studies (Bjorndal & Bolten 2008).<br />

In coral reef fishes, behaviour of larvae has been<br />

shown to influence population connectivity (Paris et al.<br />

2007) but to date, it has proven difficult to separate the<br />

role of oceanic currents from behaviour in determining<br />

distribution of oceanic juvenile marine turtles (Naro-<br />

Maciel et al. 2007). Under experimental conditions, loggerhead<br />

hatchlings have been shown to exhibit directed<br />

swimming which is consistent with that which would<br />

be needed to remain within oceanic gyres (Lohmann<br />

et al. 2001) and in the wild, oceanic juvenile loggerheads<br />

may actively select foraging areas at locations<br />

closely correlated with the latitudes of their natal beaches<br />

(Monzón-Argüello et al. 2009) and home toward<br />

their natal regions at the conclusion of the oceanic stage<br />

(Bowen et al. 2004). However, the ability to move<br />

against prevailing currents appears to be linked to<br />

increasing body size (Monzón-Argüello et al. 2009).<br />

Oceanic juvenile loggerheads in downwelling lines<br />

show minimal activity and positive buoyancy (Witherington<br />

2002) and as hawksbill hatchlings at least in<br />

some populations appear to be less active in the initial<br />

days of dispersal than other marine turtle species<br />

(Chung et al. 2009a) it is likely that small oceanic juvenile<br />

hawksbills are also relatively passive surface drifters.<br />

Therefore, while vertical movement of reef fish<br />

larvae limits the application of models based on surface<br />

current data (Fiksen et al. 2007) these models are especially<br />

applicable to marine turtles.<br />

Occurrence of active orientation, particularly at the<br />

end of the oceanic stage, may nevertheless explain<br />

contributions of hawksbill rookeries against prevailing<br />

Ó 2009 Blackwell Publishing Ltd<br />

HAWKSBILL TURTLE DISPERSAL 4849<br />

currents to foraging sites. Alternatively, unsurveyed<br />

hawksbill rookeries may contribute to foraging stocks<br />

(i.e. contributions assigned to a down-current source<br />

may instead originate from an unsurveyed or incompletely<br />

surveyed up-current source). It is also possible<br />

that a portion of Caribbean hawksbills could circle the<br />

Atlantic before re-entering the Caribbean, following a<br />

similar route to oceanic juvenile loggerheads <strong>–</strong> or perhaps<br />

a ‘short-cycle’ of the Atlantic could occasionally<br />

occur where hawksbills might become entrained in<br />

Atlantic eddies which more rapidly re-enter the Caribbean.<br />

In our comparisons between many-to-many<br />

results and the hatchling drift model, estimated contributions<br />

from some rookeries were high using genetic<br />

markers while the drift model predicted no contributions.<br />

These findings may be indicative of an Atlantic<br />

cycle or short-cycle: for instance, Mexican contributions<br />

to seemingly up-current foraging sites in the eastern<br />

Caribbean may be explained by hawksbills cycling a<br />

portion of the Atlantic before re-entering the Caribbean.<br />

Integration of better biological data with heuristic<br />

modelling of oceanic drift will assist in resolving migratory<br />

connectivity for marine turtles. Patterns of dispersal<br />

for oceanic juveniles may be confirmed by<br />

documentation of stranding patterns (Meylan & Redlow<br />

2006) and genetic studies (Carreras et al. 2006; Bowen<br />

et al. 2007a). In surveys to date, magnitude of hawksbill<br />

genetic diversity varies regionally <strong>–</strong> being lowest in the<br />

Gulf of Mexico, where mixed stock analysis indicates<br />

that oceanic juveniles originate almost exclusively from<br />

Mexican rookeries (Bowen et al. 2007a), and higher near<br />

the eastern Caribbean gyres and in foraging areas<br />

passed by the Caribbean current. Sightings and strandings<br />

of oceanic phase hawksbills in Florida (Meylan &<br />

Redlow 2006) and rarely on the east coast (Parker 1995)<br />

support transport from the Caribbean through the<br />

Florida Straits, and the presence of a neritic juvenile<br />

hawksbill foraging aggregation in Bermuda (Meylan<br />

et al. 2003) suggests some transport in the Gulf Stream.<br />

However, further genetic sampling of stranded and<br />

free-living oceanic hawksbills and genetic characterization<br />

of additional neritic foraging grounds is required.<br />

Efforts to resolve movements are particularly timely,<br />

as management of hawksbill turtles has been the subject<br />

of considerable controversy in recent years. Hawksbills<br />

are designated as ‘critically endangered’ by the IUCN<br />

(2007) and listing on Appendix I of the Convention on<br />

the <strong>International</strong> Trade in Endangered Species (CITES)<br />

bars international trade among parties to the Convention.<br />

Nevertheless, commercially valuable hawksbill<br />

products are still in high demand in some areas. Recent<br />

debate has centred on conservation status and priorities,<br />

scale and extent of transboundary movements (Bowen<br />

& Bass 1997; Mrosovsky 1997) and sustainable use of


4850 J. M. BLUMENTHAL ET AL.<br />

mixed stocks (Bowen et al. 2007b; Godfrey et al. 2007;<br />

Mortimer et al. 2007a,b). It has been argued that harvesting<br />

of hawksbills from foraging grounds should be<br />

prohibited, as exploitation of mixed stocks could potentially<br />

impact rookeries on a regional scale <strong>–</strong> or alternatively,<br />

that sustainability of harvesting on hawksbill<br />

foraging aggregations could be determined on a caseby-case<br />

basis (Godfrey et al. 2007).<br />

Many-to-many analysis and hatchling drift models<br />

facilitate an integrative, rookery-centric approach<br />

(Bolker et al. 2007) in which contributions of both small,<br />

vulnerable rookeries and large, regionally important<br />

rookeries are evaluated. Oceanic juveniles from some<br />

rookeries appear to be dispersed among multiple foraging<br />

grounds, while those from other rookeries appear to<br />

be more locally constrained. This diversity of distribution<br />

represents both a challenge and an opportunity for<br />

marine turtle management. Broadly dispersed stocks<br />

are difficult to manage, yet threats are often geographically<br />

widely dispersed and their impacts on individual<br />

stocks will depend on how they interact. In contrast, in<br />

areas with greater levels of local or proximate recruitment,<br />

threats <strong>–</strong> or conservation efforts <strong>–</strong> may have a<br />

concentrated effect (Bowen et al. 2004, 2005). Indeed,<br />

Caribbean hawksbill rookeries have experienced varying<br />

population trends, ranging from near-extirpation<br />

(Cayman Islands: Aiken et al. 2001) to dramatic increase<br />

(Antigua: Richardson et al. 2006; Barbados: Beggs et al.<br />

2007) <strong>–</strong> raising the possibility that variations in dispersal<br />

across an oceanographically complicated region<br />

may be a factor in these differing trajectories.<br />

While much remains to be determined regarding<br />

movements and management needs of hawksbill turtles,<br />

integration of many-to-many analysis and oceanographic<br />

data represents a promising approach. In this<br />

study, we delineated links between regional management<br />

units by conducting many-to-many mixed stock<br />

analysis across the range of Caribbean hawksbill rookeries<br />

and foraging aggregations, developed a tool with<br />

which to illustrate a possible role of ocean currents in<br />

determining distribution of oceanic juveniles, and highlighted<br />

gaps in knowledge and priorities for future<br />

sampling. Overall, results facilitate a broader understanding<br />

of Caribbean hawksbill movements <strong>–</strong> filling an<br />

important gap in knowledge of life-history and potentially<br />

informing regional management.<br />

Acknowledgments<br />

For invaluable logistical support and assistance with fieldwork,<br />

we thank Cayman Islands Department of Environment<br />

research, administration, operations and enforcement staff and<br />

numerous volunteers. For advice on analysis, we thank J. Hunt<br />

and W. Pitchers at University of Exeter, Cornwall campus.<br />

Vincente Guzmán (CONANP-México) and Eduardo Cuevas<br />

(Pronature-Península de Yucatán A.C.) generously provided<br />

information on population sizes of Campeche, Yucatán and<br />

Quintana Roo rookeries. Work in the Cayman Islands, the UK<br />

and the USA was generously supported by the European<br />

Social Fund, the Darwin Initiative, the Department of Environment,<br />

Food & Rural Affairs (DEFRA), Eckerd College, the Foreign<br />

and Commonwealth Office for the Overseas Territories,<br />

the Ford Foundation, the National Fish and Wildlife Foundation<br />

(NFWF) and the Turtles in the Caribbean Overseas Territories<br />

(TCOT) Project. We also acknowledge support to J.<br />

Blumenthal (University of Exeter postgraduate studentship and<br />

the Darwin Initiative). The manuscript was improved by the<br />

input of three anonymous reviewers.<br />

References<br />

Abreu-Grobois FA, Horrocks JA, Formia A et al. (2006) New<br />

mtDNA dloop primers which work for a variety of marine<br />

turtle species may increase the resolution capacity of mixed<br />

stock analyses. Poster presented at the 26th Annual<br />

Symposium on Sea Turtle Biology and Conservation, Crete,<br />

Greece, 2<strong>–</strong>8 April 2006. Available from http://www.<br />

iucnmtsg.org/genetics/meth/primers/abreu_grobois_etal_<br />

new_dloop_primers.pdf. Accessed 1 March 2008.<br />

Aiken JJ, Godley BJ, Broderick AC, Austin T, Ebanks-Petrie G,<br />

Hays GC (2001) Two hundred years after a commercial<br />

marine turtle fishery: the current status of marine turtles<br />

nesting in the Cayman Islands. Oryx, 35, 145<strong>–</strong>151.<br />

Allard MW, Miyamoto MM, Bjorndal KA, Bolten AB, Bowen<br />

BW (1994) Support for natal homing in green turtles from<br />

mitochondrial DNA sequences. Copeia, 1994, 34<strong>–</strong>41.<br />

Allen M, Engstrom AS, Meyers S et al. (1998) Mitochondrial<br />

DNA sequencing of shed hairs and saliva on robbery caps:<br />

sensitivity and matching probabilities. Journal of Forensic<br />

Science, 43, 453<strong>–</strong>464.<br />

Andersen LW, Ruzzante DE, Walton M et al. (2001)<br />

Conservation genetics of harbour porpoises, Phocoena<br />

phocoena, in eastern and central North Atlantic. Conservation<br />

Genetics, 2, 309<strong>–</strong>324.<br />

Baillie JEM, Hilton-Taylor C, Stuart SN (2004) IUCN red List of<br />

Threatened Species. A Global Species Assessment. IUCN, Gland<br />

Switzerland.<br />

Baker CS, Patenaude NJ, Bannister JL, Robins J, Kato H (1999)<br />

Distribution and diversity of mtDNA lineages among<br />

southern right whales (Eubalaena australis) from Australia<br />

and New Zealand. Marine Biology, 134, 1<strong>–</strong>7.<br />

Bass AL (1999) Genetic analysis to elucidate the natural history<br />

and behavior of hawksbill turtles (Eretmochelys imbricata) in<br />

the Wider Caribbean: a review and reanalysis. Chelonian<br />

Conservation and Biology, 3, 195<strong>–</strong>199.<br />

Bass AL, Witzell WN (2000) Demographic composition of<br />

immature green turtles (Chelonia mydas) from the east central<br />

Florida coast: evidence from mtDNA markers. Herpetologica,<br />

56, 357<strong>–</strong>367.<br />

Bass AL, Good DA, Bjorndal KA et al. (1996) Testing models<br />

of female reproductive migratory behavior and population<br />

structure in the Caribbean hawksbill turtle, Eretmochelys<br />

imbricata, with mtDNA sequences. Molecular Ecology, 5, 321<strong>–</strong><br />

328.<br />

Ó 2009 Blackwell Publishing Ltd


Bass AL, Epperly SP, Braun-McNeill J (2006) Green turtle<br />

(Chelonia mydas) foraging and nesting aggregations in the<br />

Caribbean and Atlantic: impact of currents and behavior on<br />

dispersal. Journal of Heredity, 97, 346<strong>–</strong>354.<br />

Beggs JA, Horrocks JA, Krueger BH (2007) Increase in hawksbill<br />

sea turtle Eretmochelys imbricata nesting in Barbados, West<br />

Indies. Endangered Species Research, 3, 159<strong>–</strong>168.<br />

Bell CD, Solomon JL, Blumenthal JM et al. (2007) Monitoring<br />

and conservation of critically-reduced marine turtle nesting<br />

populations: lessons from the Cayman Islands. Animal<br />

Conservation, 10, 39<strong>–</strong>47.<br />

Bjorndal KA, Bolten AB (2008) Annual variation in source<br />

contributions to a mixed stock: implications for quantifying<br />

connectivity. Molecular Ecology, <strong>17</strong>, 2185<strong>–</strong>2193.<br />

Blumenthal JM, Austin TJ, Bell CD et al. (2009a) Ecology of<br />

hawksbill turtles Eretmochelys imbricata on a western<br />

Caribbean foraging area. Chelonian Conservation and Biology,<br />

8, 1<strong>–</strong>10.<br />

Blumenthal JM, Austin TJ, Bothwell JB et al. (2009b) Diving<br />

behaviour and movements of juvenile hawksbill turtles<br />

Eretmochelys imbricata on a Caribbean coral reef. Coral Reefs,<br />

28, 55<strong>–</strong>65.<br />

Bolker BM, Okuyama T, Bjorndal KA, Bolten AB (2007)<br />

Incorporating multiple mixed stocks in mixed stock analysis:<br />

‘many-to-many’ analyses. Molecular Ecology, 16, 685<strong>–</strong>695.<br />

Bolten AB (2003) Variation in sea turtle life-history patterns:<br />

neritic vs. oceanic developmental stages. In:The Biology of Sea<br />

Turtles, Volume I (eds Lutz PL, Musick JA, Wyneken J). pp.<br />

243<strong>–</strong>257, CRC Press, Boca Raton.<br />

Bolten AB, Bjorndal KA, Martins HR et al. (1998) Trans-<br />

Atlantic developmental migrations of loggerhead sea turtles<br />

demonstrated by mtDNA sequence analyses. Ecological<br />

Applications, 8, 1<strong>–</strong>7.<br />

Botsford LW, White JW, Coffroth M-A et al. (2009)<br />

Connectivity and resilience of coral reef metapopulations in<br />

marine protected areas: matching empirical efforts to<br />

predictive needs. Coral Reefs, 28, 303<strong>–</strong>305.<br />

Bowen BW, Bass AL (1997) Movement of hawksbill turtles:<br />

what scale is relevant to conservation and what scale is<br />

resolvable with mtDNA data? Chelonian Conservation and<br />

Biology, 3, 440<strong>–</strong>442.<br />

Bowen BW, Abreu-Grobois FA, Balazs GH et al. (1995) Trans-<br />

Pacific migrations of the loggerhead sea turtle demonstrated<br />

with mitochondrial DNA markers. Proceedings of the National<br />

Academy of Sciences USA, 92, 3731<strong>–</strong>3734.<br />

Bowen BW, Bass AL, Garcia A et al. (1996) The origin of<br />

hawksbill turtles in a Caribbean feeding area as indicated by<br />

genetic markers. Ecological Applications, 6, 566<strong>–</strong>572.<br />

Bowen BW, Bass AL, Chow S et al. (2004) Natal homing in<br />

juvenile loggerhead turtles (Caretta caretta). Molecular Ecology,<br />

13, 3797<strong>–</strong>3808.<br />

Bowen BW, Bass AL, Soares L, Toonen RJ (2005) Conservation<br />

implications of complex population structure: lessons from<br />

the loggerhead turtle (Caretta caretta). Molecular Ecology, 14,<br />

2389<strong>–</strong>2402.<br />

Bowen BW, Grand WS, Hillis-Starr Z et al. (2007a) Mixed-stock<br />

analysis reveals the migrations of juvenile hawksbill turtles<br />

(Eretmochelys imbricata) in the Caribbean sea. Molecular<br />

Ecology, 16, 49<strong>–</strong>60.<br />

Bowen BW, Grant WS, Hillis-Starr Z et al. (2007b) The<br />

advocate and the scientist: debating the commercial<br />

Ó 2009 Blackwell Publishing Ltd<br />

HAWKSBILL TURTLE DISPERSAL 4851<br />

exploitation of endangered hawksbill turtles. Molecular<br />

Ecology, 16, 3514<strong>–</strong>3515.<br />

Broderick D, Moritz C, Miller JD, Guinea ML, Prince JR,<br />

Limpus CJ (1994) Genetic studies of the hawksbill turtle<br />

(Eretmochelys imbricata): evidence for multiple stocks and<br />

mixed feeding grounds in Australian waters. Pacific<br />

Conservation Biology, 1, 123<strong>–</strong>131.<br />

Browne DC, Horrocks JA, Abreu-Grobois FA (in press)<br />

Population subdivision in hawksbill turtles nesting on<br />

Barbados, West Indies, determined from mitochondrial DNA<br />

control region sequences. Conservation Genetics, DOI 10.1007/<br />

s10592-009-9883-3<br />

Carr AF (1987) New perspectives on the pelagic stage of sea<br />

turtle development. Conservation Biology, 1, 103<strong>–</strong>121.<br />

Carreras C, Pont S, Maffucci F et al. (2006) Genetic structuring<br />

of immature loggerhead sea turtles (Caretta caretta) in the<br />

Mediterranean Sea reflects water circulation patterns. Marine<br />

Biology, 149, 1269<strong>–</strong>1279.<br />

Chacón D (2004) Caribbean Hawksbills <strong>–</strong> An Introduction to<br />

Their Biology and Conservation Status. WWF-Regional<br />

Program for Latin America and the Caribbean, San Jose,<br />

Costa Rica.<br />

Chung FC, Pilcher NJ, Salmon M, Wyneken J (2009a) Offshore<br />

migratory activity of hawksbill turtle (Eretmochelys imbricata)<br />

hatchlings, I. Quantitative analysis of activity, with<br />

comparisons to green turtles (Chelonia mydas). Chelonian<br />

Conservation and Biology, 8, 28<strong>–</strong>34.<br />

Chung FC, Pilcher NJ, Salmon M, Wyneken J (2009b) Offshore<br />

migratory activity of hawksbill turtle (Eretmochelys imbricata)<br />

hatchlings II. Swimming gaits, swimming speed, and<br />

morphological comparisons. Chelonian Conservation and<br />

Biology, 8, 35<strong>–</strong>42.<br />

van Dam RP, Diez CE, Balazs GH et al. (2008) Sex-specific<br />

migration patterns of hawksbill turtles from Mona Island,<br />

Puerto Rico. Endangered Species Research, 4, 85<strong>–</strong>94.<br />

Diaz-Fernandez R, Okayama T, Uchiyama T et al. (1999)<br />

Genetic sourcing for the hawksbill turtle, Eretmochelys<br />

imbricata, in the northern Caribbean region. Chelonian<br />

Conservation and Biology, 3, 296<strong>–</strong>300.<br />

Dutton PH (1996) Methods for collection and preservation of<br />

samples for sea turtle genetic studies. In: Proceedings of the<br />

<strong>International</strong> Symposium on Sea Turtle Conservation Genetics<br />

(eds Bowen BW, Witzell WN), pp. <strong>17</strong><strong>–</strong>24. National Technical<br />

Information Service, Springfield, VA, USA, NOAA Technical<br />

Memorandum NMFS-SEFSC-396.<br />

Encalada SE, Bjorndal KA, Bolten AB et al. (1998) Population<br />

structure of loggerhead turtle (Caretta caretta) nesting<br />

colonies in the Atlantic and Mediterranean as inferred from<br />

mitochondrial DNA control region sequences. Marine<br />

Biology, 130, 567<strong>–</strong>575.<br />

Engstrom TN, Meylan PA, Meylan AB (2002) Origin of<br />

juvenile loggerhead turtles (Caretta caretta) in a tropical<br />

developmental habitat in Caribbean Panama. Animal<br />

Conservation, 5, 125<strong>–</strong>133.<br />

Excoffier L, Laval G, Schneider S (2005) Arlequin ver. 3.0: an<br />

integrated software package for population genetics data<br />

analysis. Evolutionary Bioinformatics Online, 1, 47<strong>–</strong>50.<br />

Fiksen O, Jorgensen C, Kristiansen T, Vikebo F, Huse G (2007)<br />

Linking behavioural ecology and oceanography: larval<br />

behaviour determines growth, mortality and dispersal.<br />

Marine Ecology Progress Series, 347, 195<strong>–</strong>205.


4852 J. M. BLUMENTHAL ET AL.<br />

Formia A, Godley BJ, Dontaine J-F, Bruford MW (2006)<br />

Mitochondrial DNA diversity and phylogeography of<br />

endangered green turtle (Chelonia mydas) populations in<br />

Africa. Conservation Genetics, 7, 353<strong>–</strong>369.<br />

Formia A, Broderick AC, Glen F et al. (2007) Genetic<br />

composition of the Ascension Island green turtle rookery<br />

based on mitochondrial DNA: implications for sampling and<br />

diversity. Endangered Species Research, 3, 145<strong>–</strong>158.<br />

Garduño-Andrade M (1999) Nesting of the hawksbill turtle,<br />

Eretmochelys imbricata, in Rio Lagartos, Yucatan, Mexico,<br />

1990<strong>–</strong>1997. Chelonian Conservation and Biology, 3, 281<strong>–</strong>285.<br />

Girard C, Tucker AD, Calmettes B (2009) Post-nesting<br />

migrations of loggerhead sea turtles in the Gulf of Mexico:<br />

dispersal in highly dynamic conditions. Marine Biology, 156,<br />

1827<strong>–</strong>1839.<br />

Godfrey MH, Webb GJW, Manolis SC, Mrosovsky N (2007)<br />

Hawksbill sea turtles: can phylogenetics inform harvesting?<br />

Molecular Ecology, 16, 3511<strong>–</strong>3513.<br />

Hasbún CR (2002) Observations on first day dispersal of<br />

neonatal hawksbill turtles (Eretmochelys imbricata). Marine<br />

Turtle Newsletter, 96, 7<strong>–</strong>10.<br />

Hays GC, Marsh R (1997) Estimating the age of juvenile<br />

loggerhead sea turtles in the North Atlantic. Canadian Journal<br />

of Zoology, 75, 40<strong>–</strong>46.<br />

Hood GM (2009) PopTools version 3.1.1. Available on the<br />

internet. URL http://www.cse.csiro.au/poptools. Accessed<br />

20 July 2008.<br />

Horrocks JA, Vermeer LA, Krueger B et al. (2001) Migration<br />

routes and destination characteristics of post-nesting<br />

hawksbill turtles satellite-tracked from Barbados, West<br />

Indies. Chelonian Conservation and Biology, 4, 107<strong>–</strong>114.<br />

IUCN (2007) Red List of Threatened Species. Available from<br />

http://www.iucnredlist.org. Accessed 1 April 2008.<br />

Kocher TD, Thomas WK, Meyers A et al. (1989) Dynamics of<br />

mitochondrial DNA evolution in animals: amplification and<br />

sequencing with conserved primers. Proceedings of the<br />

National Academy of Sciences of the USA, 86, 6196<strong>–</strong>6200.<br />

Lahanas PN, Bjorndal KA, Bolten AB et al. (1998) Genetic<br />

composition of a green turtle (Chelonia mydas) feeding<br />

ground population: evidence for multiple origins. Marine<br />

Biology, 130, 345<strong>–</strong>352.<br />

Laurent L, Casale P, Bradai N et al. (1998) Molecular<br />

resolution of marine turtle stock composition in fishery<br />

bycatch: a case study in the Mediterannean. Molecular<br />

Ecology, 7, 1529<strong>–</strong>1542.<br />

Lewis CB (1940) The Cayman Islands and marine turtle.<br />

In:Herpetology of the Cayman Island (ed. Grant C). pp. 56<strong>–</strong>65,<br />

Bulletin of the Institute of Jamaica Science Series, Kingston.<br />

Lohmann KJ, Cain SD, Dodge SA, Lohmann CM (2001)<br />

Regional magnetic fields as navigational markers for sea<br />

turtles. Science, 294, 364<strong>–</strong>366.<br />

Luke K, Horrocks JA, LeRoux RA, Dutton PH (2004) Origins of<br />

green turtle (Chelonia mydas) feeding aggregations around<br />

Barbados, West Indies. Marine Biology, 144, 799<strong>–</strong>805.<br />

Maffucci F, Kooistra WHCF, Bentivegna F (2006) Natal origin<br />

of loggerhead turtles, Caretta caretta, in the neritic habitat off<br />

the Italian coasts, Central Mediterranean. Biological<br />

Conservation, 127, 183<strong>–</strong>189.<br />

Meylan AB (1999) <strong>International</strong> movements of immature and<br />

adult hawksbill turtles (Eretmochelys imbricata) in the Caribbean<br />

region. Chelonian Conservation and Biology, 3, 189<strong>–</strong>194.<br />

Meylan A, Redlow A (2006) Eretmochelys imbricata <strong>–</strong> Hawksbill<br />

turtle. In: Biology and Conservation of Florida Turtles (ed.<br />

Meylan PA). Chelonian Research Foundation, Lunenburg, MA,<br />

USA, 3, 105<strong>–</strong>127.<br />

Meylan AB, Bowen BW, Avise JC (1990) A genetic test of the<br />

natal homing versus social facilitation models for green<br />

turtle migration. Science, 248, 724<strong>–</strong>727.<br />

Meylan PA, Meylan AB, Gray J, Ward J (2003) The hawksbill<br />

turtle in Bermuda. In: Proceedings of the Twenty-Second Annual<br />

Symposium on Sea Turtle Biology and Conservation (compiler,<br />

Seminoff JA), p. 26. National Technical Information Service,<br />

Springfield, VA, USA. NOAA Technical Memorandum<br />

NMFS-SEFSC-503.<br />

Millar RB (1987) Maximum likelihood estimation of mixed<br />

stock fishery composition. Canadian Journal of Fisheries and<br />

Aquatic Science, 44, 583<strong>–</strong>590.<br />

Milligan BG (1998) Total DNA isolation. In:Molecular Genetic<br />

Analysis of Populations, a Practical Approach (ed Hoelzel AR).<br />

pp. 29<strong>–</strong>64, IRL Press, Oxford.<br />

Moncada F, Carillo E, Saenz A, Nodarse G (1999)<br />

Reproduction and nesting of the hawksbill turtle,<br />

Eretmochelys imbricata, in the Cuban archipelago. Chelonian<br />

Conservation and Biology, 3, 257<strong>–</strong>263.<br />

Monzón-Argüello C, Rico C, Carreras C, Calbuig P, Marco A,<br />

López-Jurado LF (2009) Variation in spatial distribution of<br />

juvenile loggerhead turtles in the eastern Atlantic and<br />

western Mediterranean sea. Journal of Experimental Marine<br />

Biology and Ecology, 373, 79<strong>–</strong>86.<br />

Mora C, Sale PF (2002) Are populations of coral reef fish open<br />

or closed? Trends in Ecology and Evolution, <strong>17</strong>, 422<strong>–</strong>428.<br />

Mortimer JA, Donnelly M (2007) Marine Turtle Specialist<br />

Group 2007 IUCN Red List Status Assessment Hawksbill<br />

Turtle (Eretmochelys imbricata). Available from http://<br />

www.iucn-mtsg.org/red_list/ei/index.shtml. Accessed 1<br />

March 2008.<br />

Mortimer JA, Donnelly M, Meylan AB, Meylan PA (2007a)<br />

Critically endangered hawksbill turtles: molecular genetics<br />

and the broad view of recovery. Molecular Ecology, 16, 3516<strong>–</strong><br />

35<strong>17</strong>.<br />

Mortimer JA, Meylan PA, Donnelly M (2007b) Whose turtles<br />

are they, anyway? Molecular Ecology, 16, <strong>17</strong><strong>–</strong>18.<br />

Mrosovsky N (1997) Movement of hawksbill turtles <strong>–</strong> a<br />

different perspective on the DNA data. Chelonian<br />

Conservation and Biology, 3, 438<strong>–</strong>439.<br />

Musick JA, Limpus CJ (1997) Habitat utilization and migration<br />

in juvenile sea turtles. In:The Biology of Sea Turtle (eds Lutz<br />

PL, Musick JA). pp. 137<strong>–</strong>163, CRC Press, Boca Raton.<br />

Naro-Maciel E, Becker HJ, Lima EHSM, Marcovaldi MA,<br />

DeSalle R (2007) Testing dispersal hypotheses in foraging<br />

green sea turtles (Chelonia mydas) of Brazil. Journal of<br />

Heredity, 98, 29<strong>–</strong>39.<br />

Norman JA, Moritz C, Limpus CJ (1994) Mitochondrial DNA<br />

control region polymorphisms: genetic markers for<br />

ecological studies of marine turtles. Molecular Ecology, 3,<br />

363<strong>–</strong>373.<br />

Okuyama T, Bolker BM (2005) Combining genetic and<br />

ecological data to estimate sea turtle origins. Ecological<br />

Applications, 15, 315<strong>–</strong>325.<br />

Owens DW, Ruiz GW (1980) New methods of obtaining blood<br />

and cerebrospinal fluid from marine turtles. Herpetologica, 36,<br />

<strong>17</strong><strong>–</strong>20.<br />

Ó 2009 Blackwell Publishing Ltd


Paris CB, Cherubin LM, Cowen RK (2007) Surfing, spinning, or<br />

diving from reef to reef: effects on population connectivity.<br />

Marine Ecology Progress Series, 347, 285<strong>–</strong>300.<br />

Parker LG (1995) Encounter with a juvenile hawksbill turtle<br />

offshore Sapelo Island, Georgia. Marine Turtle Newsletter, 71,<br />

19<strong>–</strong>22.<br />

Pella J, Masuda M (2001) Bayesian methods for analysis of<br />

stock mixtures from genetic characters. Fishery Bulletin, 99,<br />

151<strong>–</strong>167.<br />

Pella J, Milner GB (1987) Use of genetic marks in stock<br />

composition analyses. In:Population Genetics and Fishery<br />

Managemen (eds Ryman N, Utter F). pp. 247<strong>–</strong>276, University<br />

of Washington Press, Seattle.<br />

Reece JS, Ehrhart LM, Parkinson CL (2006) Mixed stock<br />

analysis of juvenile loggerheads (Caretta caretta) in Indian<br />

River lagoon, Florida: implications for conservation<br />

planning. Conservation Genetics, 7, 345<strong>–</strong>352.<br />

Richardson JI, Hall DB, Mason PA et al. (2006) Eighteen years<br />

of saturation tagging data reveal a significant increase in<br />

nesting hawksbill sea turtles (Eretmochelys imbricata) on Long<br />

Island, Antigua. Animal Conservation, 9, 302<strong>–</strong>307.<br />

Roberts H (1994) Reefs and lagoons of Grand Cayman. In:The<br />

Cayman Islands: Natural History and Biogeograph (eds Brunt<br />

MA, Davies JE). pp. 75<strong>–</strong>104, Kluwer Academic Publishers,<br />

Netherlands.<br />

Spalding MD, Fox HE, Allen GR et al. (2007) Marine<br />

ecoregions of the world: a bioregionalization of coast and<br />

shelf areas. BioScience, 57, 573<strong>–</strong>583.<br />

Troëng S, Dutton PH, Evans D (2005) Migration of hawksbill<br />

turtles Eretmochelys imbricata from Tortuguero, Costa Rica.<br />

Ecography, 28, 394<strong>–</strong>402.<br />

Velez-Zuazo X, Ramos WD, van Dam RP et al. (2008) Dispersal,<br />

recruitment and migratory behaviour in a hawksbill sea turtle<br />

aggregation. Molecular Ecology, <strong>17</strong>, 839<strong>–</strong>853.<br />

Wessel P, Smith HF (1991) Free software helps map and<br />

display data. Eos, Transactions of the American Geophysical<br />

Union, 72, 441.<br />

Ó 2009 Blackwell Publishing Ltd<br />

HAWKSBILL TURTLE DISPERSAL 4853<br />

Whiting SD, Koch AU (2006) Oceanic movement of a benthic<br />

foraging juvenile hawksbill turtle from the Cocos (Keeling)<br />

Islands. Marine Turtle Newsletter, 112, 15<strong>–</strong>16.<br />

Witherington BE (2002) Ecology of neonate loggerhead turtles<br />

inhabiting lines of downwelling near a Gulf Stream front.<br />

Marine Biology, 140, 843<strong>–</strong>853.<br />

Witteveen BH, Straley JM, von Ziegesar O, Steel D, Baker CS<br />

(2004) Abundance and mtDNA differentiation of humpback<br />

whales (Megaptera novaeangliae) in the Shumagin Islands,<br />

Alaska. Canadian Journal of Zoology, 82, 1352<strong>–</strong>1359.<br />

J.M.B. is a research officer at the Cayman Islands Department<br />

of Environment, where her work centres on using spatial ecology<br />

to aid in understanding management requirements of sea<br />

turtles and other marine species. G.E.-P. is the director of the<br />

Department of Environment and T.J.A. is the deputy director<br />

for the research section. M.S.C. coordinates a gobal sea turtle<br />

network as director of seaturtle.org. F.A.A.-G.’s research on<br />

conservation genetics spans marine turtle populations in Mexican<br />

and Wider Caribbean sites. P.A.M. is RR Hallin Professor<br />

of Natural Science at Eckerd College; A.B.M. leads the marine<br />

turtle research program for the Florida Fish and Wildlife Conservation<br />

Commission; together they study the reproductive<br />

biology, ecology and migrations of sea turtles in the Western<br />

Atlantic with long-term projects in Panama and Bermuda. A.F.<br />

works on marine turtle conservation genetics in West Africa<br />

and coordinates the Marine Turtle Partnership in Gabon.<br />

M.W.B. is head of the Biodiversity and Ecological Processes<br />

Group at Cardiff University. A.C.B. and B.J.G. coordinate the<br />

Marine Turtle Research Group at a range of sites around the<br />

world including Ascension Island, Northern Cyprus and the<br />

UK Overseas Territories.


The Elusive Manatee --<br />

An ethological approach to understanding behavior in the West Indian manatee<br />

Caryn Self Sullivan, Department of Wildlife and Fisheries Sciences, Texas A&M University, College Station, TX 77843-2258<br />

Last updated 19 October 2000<br />

CONTENTS<br />

Preface<br />

Part I: Introduction --<br />

Ethology, Proximate & Ultimate, and <strong>Sirenian</strong>s<br />

Part II: Problem Solving --<br />

Social, Reproductive & Physical<br />

References<br />

Recommended Reading<br />

Glossary<br />

Acknowledgments<br />

PREFACE<br />

This Species Brief originated in WFSC 422, an<br />

ethology course taught by my academic advisor, Dr. Jane<br />

M. Packard, at Texas A&M University in 1998. It was<br />

updated this year for distribution to Earthwatch Institute<br />

volunteers. Please consider it a work in progress <strong>–</strong> a “draft<br />

document” as it is continuously being revised and updated<br />

with better references and new information. Designed to be<br />

used by schools, zoos, wildlife parks, and oceanaria, it<br />

makes an excellent starting place for students, teachers, and<br />

others who are interested in learning about animal behavior<br />

and/or sirenians (manatees & dugongs). But, remember, it<br />

is only a briefing document. Use it to catapult yourself into<br />

the exciting world of animal behavior -- using manatees as<br />

an example. For more details, start with the<br />

Recommended Reading section; scientists and university<br />

students are encouraged to delve into the primary literature<br />

listed in the References section.<br />

PART I: INTRODUCTION<br />

As we idled around the corner of Swallow Caye I<br />

sighted two manatee noses in the distance. They were<br />

barely visible as they broke the surface of the clear<br />

Caribbean water. Patch, our boat operator, spotted the<br />

manatees at almost the same instant <strong>–</strong> he probably saw<br />

them first -- because before I could motion to him, he had<br />

already shut down the engine. We waited in silence, hoping<br />

they would surface again. So it goes with research on the<br />

elusive manatee. Most behavioral observations of manatees<br />

have been conducted on Florida manatees, either in<br />

captivity or in the clear spring waters of Florida during<br />

winter aggregations. Only recently have we attempted to<br />

observe behavior in Antillean manatees, which are sparsely<br />

distributed throughout the Caribbean, including the tropical<br />

waters of Belize.<br />

Five minutes passed - how long can these guys stay<br />

down? What are they doing down there? One reason we<br />

know so little about these incredibly well adapted animals<br />

is because they spend the majority of their time underwater,<br />

1<br />

regularly staying submerged 3-5 minutes between breaths.<br />

We heard them before we saw them. Both noses broke the<br />

water with a forceful exhalation at virtually the same<br />

moment. Then they were down again. I quietly entered the<br />

water and stealthily snorkeled the 50 meters towards their<br />

last location. Where did they go? Stop. Look. Listen. I<br />

heard them breathe again. When they finally came into<br />

view underwater, I thought, "Uh oh... a mother calf pair --<br />

they are going to run away".<br />

The larger animal was about 3 meters long, almost<br />

twice as big as the smaller. Readings and previous<br />

experience led me to assume a mother-calf relationship<br />

based on this size differential. But they didn't run. The<br />

larger animal was gently nuzzling the smaller one's back<br />

with its big prehensile lips. The next time they surfaced to<br />

breath, they were nose to nose in a manatee "kiss". When<br />

they noticed us, the nuzzling stopped and they both sank<br />

slowly to the soft muddy bottom. As they sank, I heard a<br />

few squeaks, similar to manatee vocalizations I’d recorded<br />

between mother-calf pairs last year. There they rested, side<br />

by side in typical mother-calf position, for three minutes.<br />

When the smaller one rose to the surface to breathe, I could<br />

tell it was a female by the location of a genital slit near the<br />

anus. But, what a surprise I had when the larger animal<br />

surfaced and I saw by a genital slit near the umbilicus scar<br />

that it was a male. I'll make no more assumptions about<br />

mother-calf pairs based on size differentials!<br />

Ethology<br />

Do manatees breath simultaneously? If so, why?<br />

Why was the large male manatee nuzzling the smaller<br />

female? Why do manatees "kiss"? How did they sink to<br />

the bottom and stay - without moving a muscle? Why did<br />

they vocalize during their descent? How do manatees<br />

create sound? Do manatees often lie side by side on the<br />

bottom? Are manatees usually found in pairs, or groups, or<br />

alone? Why? How long can manatees stay underwater<br />

without breathing? Why did the smaller animal surface to<br />

breath before the larger animal did? These are just a few of<br />

the questions raised by the brief observation. Some<br />

answers to these "how" and "why" questions are known;<br />

other answers may come through long-term ethological<br />

studies.<br />

Ethology is a relatively new, multi-perspective<br />

scientific approach to the study of animal behavior. Made<br />

famous by the work of 1973 Nobel Prize winners, Konrad<br />

Lorenz, Karl von Frisch, and Nikolaas Tinbergen<br />

(www.nobel.se/medicine/laureates/1973/index.html), it<br />

focuses on animal behavior in a natural setting. By using a<br />

Scientific Perspective, it differs from Folk Psychology,<br />

which is often used to explain animal behavior to the


general public. Folk psychology perspectives are intuitive<br />

in nature, usually based on personal experiences and<br />

observations. They are considered anthropomorphic<br />

because they describe and explain behavior in human terms<br />

<strong>–</strong> which are often the only terms we have to start with!<br />

These perspectives are appropriate and very useful when<br />

communicating with non-scientists, such as audiences in<br />

zoos, oceanaria, and wildlife parks. An interpreter will<br />

often use folk psychology to describe and explain animal<br />

behaviors based on the “model” that animals have desires,<br />

beliefs, and emotions like humans. Scientists also use these<br />

perspectives in developing hypotheses about specific<br />

behaviors. For example, I was using folk psychology when<br />

I assumed that the manatees in the anecdote above were a<br />

mother-calf pair. My intuition was based on the size<br />

differential and behavioral patterns typical of mother-calf<br />

pairs.<br />

Ethology encourages us to develop additional<br />

Scientific Perspectives in understanding, explaining,<br />

and/or describing animal behavior; the classical ethological<br />

perspectives include cause, development, evolution, and<br />

function (Martin & Bateson 1993, Lehner 1996). Modern<br />

ethologists agree that the behavior of an animal is the result<br />

of complex interactions between the genetic makeup of an<br />

individual and environmental factors that act upon the<br />

individual (Alcock 1998). However, many aspects of an<br />

animal's behavior can be explained from two very different<br />

perspectives: proximate and ultimate (Martin and Bateson<br />

1993, Lehner 1996). This often results in<br />

miscommunication among observers who are looking at<br />

behavior from different perspectives. Proximate and<br />

ultimate comparisons are equivalent to apple and orange<br />

comparisons <strong>–</strong> i.e. they are both valid fruits, but they are<br />

different things. "How" questions are usually asked from a<br />

proximate perspective; how questions seek explanations<br />

about the physical and chemical mechanisms that trigger an<br />

individual animal's behavior at any given point in time.<br />

"Why" questions, on the other hand, are usually based on<br />

ultimate perspectives; answers to these questions attempt<br />

to explain why certain behaviors exist within a population<br />

(or species) of animals. In other words, what pressures of<br />

natural selection led to the existence of a particular<br />

behavior within a population or species. Ethologists<br />

further divide proximate and ultimate into the subcategories<br />

of cause, development, evolution, and function<br />

based on the work of Niko Tinbergen (Martin and Bateson<br />

1993, Lehner 1996).<br />

Proximate: Cause and Development are<br />

proximate perspectives, which look at the behavior of an<br />

individual animal. Proximate Cause perspectives include<br />

looking at both internal mechanisms (hormones,<br />

neurotransmitters) and external stimuli (pheromones, photoperiod,<br />

temperature) that interact to trigger specific<br />

behaviors in a mature animal. Dr. Jane M. Packard<br />

explains it using the analogy of a camera, “Think of<br />

Proximate Cause as a snapshot in time that shows what is<br />

causing the behavior at that particular moment.” For<br />

2<br />

example, in our observation above, what “caused” the<br />

manatees to kiss? Was the female was giving off some<br />

signal (vocal, chemical, or behavioral) that attracted the<br />

male? Did the tactile stimulation by the male cause<br />

hormone production in the female, which triggered the<br />

“kiss”? Most likely, it was is a complex interaction<br />

between both the internal state of each animal and the<br />

resulting external behavioral stimuli. Proximate<br />

Development perspectives look at behavioral changes that<br />

occur as an individual animal matures. Think of Proximate<br />

Development as a "video" that shows how a behavior<br />

develops and changes over time as an individual animal<br />

matures. How might the behavior of manatees at different<br />

ages compare to the interactions we observed?<br />

Ultimate: Evolution and Function are ultimate<br />

perspectives, which look at specific behaviors present in a<br />

population of animals. These behaviors are thought to<br />

have evolved over time through the process called Natural<br />

selection. For Natural selection to act on a behavioral<br />

characteristic, the behavior must meet certain criteria <strong>–</strong> the<br />

same criteria necessary for natural selection to act on a<br />

physical trait such as coloration: (1) the trait must vary<br />

among individuals within a population; (2) the variation<br />

must be heritable; (3) if the heritable variation results in<br />

differential fitness (i.e. variations in the trait result in some<br />

individuals reproducing more successfully than others);<br />

then (4) we would expect the behavior to become<br />

genetically fixed in the population as the proportion of<br />

individuals displaying the trait increased (i.e. changes in<br />

the proportion of genotype and resulting phenotype).<br />

Ultimate Evolution perspectives include the<br />

comparison of behaviors among different, closely related<br />

species. This is our “video” perspective. From an<br />

Ultimate Evolution perspective, we hypothesize about how<br />

a behavior has changed (or remained the same) at the<br />

population and/or the species level over many generations.<br />

In my study of Antillean manatees, I will be comparing<br />

behavior to previous observations of behavior in Florida<br />

manatees. Although these sub-species are very closely<br />

related, they share different habitats. The Florida manatee<br />

inhabits a temporal/sub-tropic region where its behaviors<br />

are shaped by dramatic changes in water temperature during<br />

the year. On the other hand, the Antillean manatee inhabits<br />

a tropical region where the water temperature is relatively<br />

constant year round. We expect some behaviors to differ<br />

between the two sub-species as they evolved in different<br />

habitats. If behaviors, which we think are driven by water<br />

temperature today, exist in both sub-species, then perhaps<br />

they had some other function in the past.<br />

Ultimate Function perspectives attempt to explain<br />

what the function of a specific behavior is within a<br />

population, (i.e. why animals that display this behavioral<br />

trait are more reproductively successful than individuals<br />

who do not). This is our “snapshot” perspective. If<br />

variation exists within the behavior, Ultimate Function is<br />

the perspective used to explain why. Some Florida<br />

manatees travel long distances into more temperate regions


during the summer months while others stay in the same<br />

area year round <strong>–</strong> but we are not sure why the traveling<br />

behavior exists. The traveling animal must use more<br />

energy than the year round resident. Perhaps male animals<br />

that travel are exposed to more potential mates <strong>–</strong> thereby<br />

increasing their reproductive success. Chessie, a famous<br />

male manatee first sighted in the Chesapeake Bay in 1994,<br />

is known to have traveled between Florida in the winter and<br />

Rhode Island in the summer of 1995! Sweet Pea, a female<br />

rescued near Houston, Texas, later traveled along both<br />

coasts of Florida. Gina, a female manatee first sighted in<br />

Tampa Bay on the west coast of Florida is currently<br />

hanging out in the Bahamas! One of my questions is: Do<br />

Antillean manatees exhibit similar long distance traveling<br />

behaviors?<br />

Hypotheses about “why” these behaviors exist in<br />

manatees are different from hypotheses about “how” these<br />

behaviors are executed. The “why” questions are from<br />

ultimate perspectives of evolution and function, the “how”<br />

questions are from proximate perspectives of cause and<br />

development. These four basic concepts of ethology can<br />

be arranged in a 2 x 2 table comparing TIME FRAME and<br />

ANALYSIS PERSPECTIVES:<br />

TIME/<br />

ANALYSIS<br />

Proximate<br />

Perspective<br />

Individual Animals<br />

“How Questions”<br />

Ultimate<br />

Perspective<br />

Populations/Species<br />

“Why Questions”<br />

Pattern-Static<br />

“Snapshot”<br />

CAUSE<br />

(control)<br />

behavioral<br />

triggers:<br />

internal state/<br />

external<br />

stimuli<br />

FUNCTION<br />

adaptive<br />

significance:<br />

effect on<br />

reproductive<br />

fitness<br />

Process-Dynamic<br />

“Video”<br />

DEVELOPMENT<br />

(ontogeny)<br />

changes in<br />

behavior as an<br />

animal ages:<br />

maturation and/or<br />

learning<br />

EVOLUTION<br />

(phylogeny)<br />

changes in<br />

behavior<br />

(genotype) as<br />

populations/<br />

species diverge<br />

We can remember the concepts of ethology with the<br />

acronym AB=CDEF (Animal Behavior = Cause,<br />

Development, Evolution, Function). On the TIME axis,<br />

Cause and Function are “snapshot” perspectives that look<br />

at internal state and external stimuli in an individual animal<br />

or at reproductive success in a population of animals.<br />

Development and Evolution are “video” perspectives that<br />

look at changes in behaviors over time, either as the<br />

individual animal matures or as the population evolves.<br />

On the ANALYSIS axis, Cause and Development are<br />

proximate perspectives that attempt to answer "how"<br />

questions at the level of individual animals. Evolution and<br />

Function are ultimate perspectives that attempt to answer<br />

"why" questions at the level of populations and/or species.<br />

3<br />

For more information on Ethology, I encourage you to visit<br />

Dr. Packard’s website at canis.tamu.edu/wfscCourses/.<br />

<strong>Sirenian</strong>s<br />

So what are manatees anyway, and why should we<br />

study their behavior? Manatees belong to the order Sirenia<br />

of which there are only 4 extant species in 2 families,<br />

Trichechidae and Dugongidae. Although scientists often<br />

lump sirenians together with the order Cetacea (whales<br />

and dolphins) as totally aquatic marine mammals, manatees<br />

and the dugong are actually more closely related to<br />

elephants, hyraxes, and aardvarks than to any other marine<br />

mammal (Fischer 1990, Maluf 1995, Springer et al. 1997,<br />

Gaeth et al. 1999). Until a few years ago, very few people<br />

had ever heard of manatees, dugongs, or sea cows. But, as<br />

we learn more about these elusive and highly specialized<br />

creatures that share our coastal habitats, they are becoming<br />

more and more popular among both scientists and<br />

conservationists. The West Indian manatee (Trichechus<br />

manatus), the West African manatee (Trichechus<br />

senegalensis), and the Amazonian manatee (Trichechus<br />

inunguis) are members of the family Trichechidae. The<br />

dugong (Dugong dugon) is the only surviving member of<br />

the family Dugongidae (Reynolds and Odell 1991).<br />

Steller's sea cow (Hydrodamalis gigas) is usually included<br />

when we talk about modern sirenians; it was in the family<br />

Dugongidae (Reynolds and Odell 1991), but the species<br />

was extirpated by humans in <strong>17</strong>68, just 27 years after it<br />

was discovered by Russian explorers in <strong>17</strong>41 (Stejneger<br />

1887). Today, local and international laws protect all four<br />

living species, but they are also either threatened or<br />

endangered by humans wherever they exist.<br />

Florida Manatees: There is little evidence that<br />

Florida manatees were ever harvested commercially. But<br />

subsistence use, habitat destruction, and competition for<br />

space with recreational boaters have taken their toll on both<br />

prehistoric and modern populations. The USGS Sirenia<br />

Project, the U. S. Fish and Wildlife Service, and the Florida<br />

Fish and Wildlife Conservation Commission (formerly the<br />

Florida Department of Environmental Protection) have<br />

funded much of the manatee research in the United States.<br />

Over the past three decades, conservation efforts in Florida<br />

resulted in significant scientific research on the distribution,<br />

population biology, and behavior of the Florida manatee<br />

subspecies, T. m. latirostris (see O'Shea et al. 1995).<br />

Ecological and behavioral studies on localized populations<br />

such as those in Crystal and Homosassa Rivers (Hartman<br />

1979), St. John's River (Bengtson 1981), and Sarasota Bay<br />

(Koelsch 1997) have added to our understanding of Florida<br />

manatee behavior. However, relatively little research has<br />

been conducted outside of Florida. Therefore, most of the<br />

referenced information contained herein refers to the<br />

Florida subspecies.<br />

Antillean Manatees: Even before Russian sailors<br />

were exploiting Steller’s sea cow meat in the North Pacific,<br />

European explorers were provisioning their ships with<br />

Antillean manatee meat from the Caribbean area. Besides<br />

harvesting manatees for subsistence, some Native


Americans also sold manatee meat to the Europeans<br />

(O’Shea 1994). Today, the Antillean subspecies, T. m.<br />

manatus, is classified as endangered throughout its sparse<br />

distribution in the Caribbean Sea and Western Tropical<br />

Atlantic Ocean (Lefebvre et al. 1989). O'Shea and<br />

Salisbury (1991) suggest that Belize (formerly British<br />

Honduras), where manatees have been protected since the<br />

1930s (Auil 1998), may be the last stronghold for Antillean<br />

manatees in the Caribbean. Current research in Belize by<br />

James A. "Buddy" Powell, Nicole Auil, Greg Smith, Katie<br />

LaCommare, and myself is expanding our knowledge of the<br />

Antillean manatee. Most of the anecdotes included in this<br />

brief come from my personal experiences in Belize.<br />

PART II: PROBLEM-SOLVING<br />

One way to interpret animal behavior is as a method<br />

of problem solving. Over geological time, animals have<br />

evolved behaviors that enable them to solve problems.<br />

Some of these behaviors are identical from one individual<br />

to another. We describe such a behavior as a Fixed Action<br />

Pattern (FAP), because the individual’s genes control the<br />

trait. In other words, the genetic trait has become fixed in<br />

the population and all individuals perform the behavior in<br />

exactly the same way because they inherited the trait from<br />

their ancestors. At the other end of the behavioral scale, we<br />

find behaviors that vary a great deal among individuals.<br />

We describe such a behavior as a Variable Action Pattern<br />

(VAP), because the environment controls the trait. In other<br />

words, each individual performs the behavior differently<br />

due to different environmental factors during development.<br />

Additionally, any individual may perform the behavior<br />

differently at different times, depending its internal state<br />

(hormones, chemicals, neurons) and on external stimuli<br />

(environment). Of course FAP and VAP are not specific<br />

categories, but are the end points along a continuum. If a<br />

behavior falls near the middle of this continuum, we<br />

describe it as a Modal Action Pattern (MAP). In other<br />

words, the behavior is controlled in part by genetics and in<br />

part by the environment.<br />

We can divide problem solving into three major<br />

categories: reproductive, physical, and social. Manatees<br />

have evolved some interesting behaviors to overcome<br />

reproductive and physical problems. But, they are not<br />

considered to be very social animals. Although they tend<br />

to aggregate on resources, they do not appear to live in<br />

social groups and significant social behaviors have not been<br />

observed outside of reproductive activities. This species<br />

brief will use the concepts of ethology to introduce you to<br />

the behavioral methods manatees use to solve some of their<br />

reproductive and physical problems. We will examine<br />

specific behaviors using the different perspectives of<br />

proximate cause, proximate development, ultimate<br />

evolution, and ultimate function to answer a few questions<br />

regarding "how" and "why" manatees behave as the do.<br />

4<br />

Reproductive Problem Solving<br />

We idled into one of my favorite coves at the end of<br />

Bogue C in the Drowned Cayes near Belize City. I had<br />

taken volunteer researchers to this spot on previous<br />

occasions and ALWAYS, there had been a manatee resting<br />

in the manatee hole on the far side of the cove. As if on<br />

cue, before we could even cut the engine and anchor the<br />

boat, a single manatee surfaced in the vicinity of the<br />

manatee hole. Within minutes a second manatee surfaced<br />

in the middle of the cove. Two minutes later, a third animal<br />

entered the cove. "Gee...” I thought, "This must be a<br />

popular resting cove!" But, the two animals in the center of<br />

the cove were too active to be resting. I didn’t think they<br />

could be feeding, either, because previous habitat snorkels<br />

had found NO vegetation on the bottom. Another four<br />

minutes passed and two more manatees swam under the<br />

boat to join the active pair in the middle of the cove. "This<br />

is great,” I told the volunteers, "we'll be able to see how<br />

long it takes them to settle into a resting pattern". But they<br />

didn't settle. For the next hour we watched the four<br />

manatees in the middle of the cove breathe, roll, dive, and<br />

kiss while the first animal appeared oblivious to all the<br />

activity less than 50 meters away. Were we observing a<br />

mating herd?<br />

Mating System: One parameter of the West Indian<br />

manatee mating system is known as the mating herd. A<br />

mating system is the species-typical pattern of problem<br />

solving that includes how an individual finds a mate, how<br />

long it remains with the mate, and how much energy it<br />

invests in its offspring (Drickamer, et al. 1996). The West<br />

Indian manatee mating system can be broadly defined as<br />

promiscuous with the estrous female exhibiting<br />

polyandrous behavior and the male exhibiting polygynous<br />

behavior (Hartman 1979). A manatee-mating herd<br />

consists of a group of males in pursuit of an estrous female.<br />

The group is ephemeral, lasting only from a week to a<br />

month (Hartman 1979) and consisting of up to 20 males.<br />

The group does not remain together afterwards. Males will<br />

participate in multiple mating herds and attempt to<br />

copulate with many estrous females; similarly, females will<br />

copulate with multiple males among those in the herd.<br />

When we discuss why this mating system exists in<br />

manatees, we are using the ultimate function perspective.<br />

From the perspective of proximate cause, we do not<br />

know exactly what external signal stimulates males to<br />

aggregate around and attempt to copulate with the estrous<br />

female. Females must produce some sort of signal,<br />

possibly a chemical or acoustical signal, which stimulates<br />

an internal hormonal mechanism in males causing them to<br />

pursue her. Likewise, the male’s internal state must be such<br />

that he responds to the signal. Proximate development<br />

perspectives would look at how the males’ reaction to such<br />

a signal might differ at other stages of maturity.<br />

Daniel S. Hartman, one of the first scientists to<br />

make long-term observations of manatee behavior in the<br />

wild, found similarities between manatee mating herds and<br />

elephant mating, noting that female elephants are also


polyandrous - often mating with several males over a<br />

period of several hours. When Hartman (1979) compares<br />

the mating behavior of manatees to that of elephants, he is<br />

writing from an ultimate evolution perspective. In other<br />

words, he is hypothesizing that this aspect of the mating<br />

system evolved millions of years ago in an ancestor shared<br />

by both the manatee and the elephant. Since sirenians and<br />

proboscideans are two of only four extant orders that share<br />

a common ancestor among them, manatees are often<br />

compared to elephants using the ultimate evolution<br />

perspective [NOTE: The other two orders contain the<br />

hyraxes and the aardvarks, which are rarely compared to<br />

manatees in the literature]. Another promiscuous aspect of<br />

the manatee mating system is scramble-polygyny, where<br />

multiple males attempt to mate with the estrous female, but<br />

- without overt competition. Although males aggregate on<br />

the estrous female and jockey for the best position - they<br />

exhibit little agonistic behavior. Interestingly, male<br />

dugongs appear to be more agonistic during mating events.<br />

They set up territories and exhibit lek mating behaviors<br />

(Anderson 1997). What perspective would we use to<br />

compare mating strategies between manatees and dugongs?<br />

Timing: Let's assume that our observation was of a<br />

mating herd. That is, the group of manatees in the center<br />

of the cove consisted of 1 estrous female and 3 males.<br />

Why was the first manatee, the one originally sighted in the<br />

resting hole, not involved in the mating herd? Looking at<br />

the situation from a proximate perspective, there are<br />

several possibilities, and all involve timing. Suppose the<br />

resting manatee was a female. If she was sexually mature,<br />

but not in estrous, the mating herd would have no interest<br />

as she would not be producing an estrous signal. An<br />

estrous signal is the proximate cause of the mating herd<br />

behavior. It's "how" the males know the female is ready to<br />

conceive. Similarly, if the female were sexually immature,<br />

she could not be in estrous and therefore would not be<br />

sending a signal. Scientists have only recently answered<br />

the question of when a female manatee becomes sexually<br />

mature, thanks to the development of a new aging technique<br />

by Miriam Marmontel, et al. (1990). Since manatees<br />

continuously regenerate new teeth throughout their lives<br />

(Domning and Hayek 1984), they cannot be aged by their<br />

dentition like many other marine mammals. But, by<br />

looking at growth layers in manatee ear bones, we are now<br />

reasonably confident that female manatees in Florida reach<br />

sexual maturity between the age of 3 and 4 years - most<br />

giving birth to their first calf at age 4 (Marmontel 1995).<br />

Questions of "how" the behavior of signaling develops in<br />

females as they mature fall under the proximate<br />

development perspective.<br />

On the other hand, if the resting manatee was a<br />

male, why wasn't he attracted to the estrous female in the<br />

middle of the cove? He could have been either sexually<br />

immature or sexually inactive. Using the presence or<br />

absence of sperm in the testes as an indicator, Hernandez et<br />

al. (1995) found that sexual maturity (proximate<br />

development) varied among Florida male manatees with<br />

5<br />

both size and age with some males becoming<br />

physiologically mature as young as 2 years and as small as<br />

237 cm. But, from a proximate cause perspective, they<br />

also found that the reproductive system varied in<br />

functionality among mature male manatees depending on<br />

season in Florida, with little evidence of spermatogenesis<br />

present during winter months (December <strong>–</strong> February). In<br />

many mammals, reproductive activity varies seasonally<br />

with photoperiod, or the number of light hours per day.<br />

The pineal gland is usually the organ associated with<br />

behavioral changes affected by photoperiod. However, no<br />

pineal gland has ever been found in manatees or dugongs<br />

(Ralph et al. 1985, W. Welker personal communication<br />

2000). From an ultimate evolution perspective, it is<br />

interesting that the literature is unclear regarding the<br />

existence of a pineal gland in elephants (Ralph et al. 1985).<br />

Whether the resting manatee was inactive or<br />

immature, his timing would have been out of sync with the<br />

female and the estrous signal would have no effect on his<br />

behavior. From an ultimate function perspective, we say<br />

that those manatees whose sexual behavior is triggered at<br />

the appropriate time (i.e. when both the male and female are<br />

sexually mature, active, and receptive) are more<br />

reproductively successful than manatees that waste energy<br />

on futile sexual encounters. From an ultimate evolution<br />

perspective, there have been some behaviors observed in<br />

Antillean manatees that might be associated with seasonal<br />

spermatogenesis (G. Smith unpublished data). More<br />

studies are necessary before we can determine if seasonality<br />

affects the reproductive behavior of manatees in Belize.<br />

Parental Care: The final aspect of reproductive<br />

problem solving discussed here is parental care. Like many<br />

mammals, female manatees invest considerable time and<br />

energy into a relatively small number of offspring as their<br />

reproductive strategy. From an ultimate function<br />

perspective, data collected by scientists in Florida suggest<br />

that those females that invest 2 years of parental care in<br />

each offspring prior to becoming pregnant again are more<br />

successful than other females (Marmontel 1995). Although<br />

calves begin eating on their own within 3 months of birth,<br />

they continue to nurse periodically (Hartman 1979) as they<br />

grow and learn migration routes from their mothers (R.<br />

Bonde, personal communication 1999). When we ask<br />

"why" this behavior exists, we are asking ultimate function<br />

questions. Perhaps calves need the extra protein and fat<br />

provided by mother's milk during developmental years. Or,<br />

perhaps it takes calves almost two years to learn the routes<br />

to warm water effluents and good foraging grounds<br />

necessary for survival through the temperate winters in<br />

northern and central Florida.<br />

In an ongoing study of Antillean manatees in the<br />

Southern Lagoon of Belize, Buddy Powell is also seeing<br />

mother-calf pairs remain together for long periods of time<br />

(www.wesave.org/manatee/). When we compare this<br />

behavior between the Florida and Antillean subspecies, we<br />

are using the ultimate evolutionary perspective. On the<br />

other hand, "how" the mother-calf pair remains together


during this period is a proximate cause question. It<br />

appears that calves remain in close association with their<br />

mothers via vocalizations (Hartman 1979; Reynolds 1981;<br />

Bengtson and Fitzgerald 1985; personal observations).<br />

Although manatees usually have only one calf at a<br />

time, there are rare occurrences of manatees giving birth to<br />

twins (Hartman 1979; O'Shea and Hartley 1995; Rathbun et<br />

al. 1995). Twinning is often followed by the death of one<br />

(O'Shea and Hartley 1995) or both (L. Lefebvre, personal<br />

communication; personal observation) offspring. From an<br />

ultimate function perspective, this alternative behavior<br />

raises the questions (1)"why" aren't females that have twins<br />

more successful than those that have singles?" and (2) "why<br />

has the variation (single vs. twins) in birthing behavior<br />

persisted within the West Indian manatee?" From an<br />

ultimate evolution perspective, the variation could be<br />

studied by looking at closely related species.<br />

Unfortunately, there are little data available on the<br />

occurrence of twins within other manatee species, but<br />

Marsh (1995) references vague reports of twin fetuses in<br />

dugongs. This is one line of evidence that twinning is a<br />

characteristic shared with other sirenians and not recently<br />

derived within the Florida population.<br />

Before we leave parental care, we should ask, "could<br />

the observation described in the introduction have been a<br />

manatee father caring for his offspring?" Probably not,<br />

there is no evidence that male manatees participate in any<br />

form of parental care. A better hypothesis, based on what<br />

we now know about manatee reproductive behavior, is that<br />

perhaps the large male was attracted to little female because<br />

she was sending an estrous signal.<br />

Summary: We have learned much about Florida<br />

manatee reproductive strategies over the past three decades.<br />

Studies on free-ranging manatee populations at warm<br />

water effluents and data collected from carcasses through<br />

the salvage network agree on many life history traits<br />

(O’Shea et al. 1995). Female manatees appear to reach<br />

sexual maturity at age 3, producing their first calf at age 4.<br />

Single births are the norm with rare cases of twins. More<br />

twins are reported in carcasses than observed in live<br />

manatees, leading to the assumption that uniparity is the<br />

more successful behavior. Gestation is about 1 year and<br />

calves stay with their mothers for about 2 years, making the<br />

minimum interval between successful reproductions about 3<br />

years. However, females who abort fetuses or lose a young<br />

calf may reproduce again sooner. Florida males exhibit<br />

seasonal spermatogenesis, which correlates with seasonal<br />

observations of females with neonate calves.<br />

PHYSICAL PROBLEM-SOLVING<br />

<strong>Sirenian</strong>s (manatees and dugongs) belong to group<br />

of animals commonly referred to as marine mammals.<br />

Other marine mammals include whales and dolphins; seals,<br />

sea lions, and walruses; sea otters; and polar bears.<br />

Although they are not closely related to each other<br />

(remember, manatees are more closely related to elephants<br />

than to other marine mammals), these groups share<br />

6<br />

convergent characteristics that evolved as they solved the<br />

physical problems associated with adaptation from a<br />

terrestrial to a marine environment. For example, all<br />

marine mammals must breathe air, and they have evolved in<br />

various ways that enable them to survive in an aquatic<br />

environment. In the dolphins, nostrils have migrated up the<br />

rostrum to the top of the head and become a single<br />

blowhole. Many large whales and seals have physiological<br />

adaptations that enable them to remain underwater for hours<br />

at a time.<br />

From an ultimate evolutionary perspective, one<br />

interesting hypothesis is that the common ancestor between<br />

manatees and elephants was an aquatic (rather than a<br />

terrestrial) mammal. If true, this would make the elephant<br />

the only known animal to move from the sea to the land (as<br />

all mammal ancestors did when their remote ancestors<br />

evolved from fish to amphibians), back to the sea (as did<br />

the cetaceans, pinnipeds, and sirenians) and then back to<br />

the land again. This idea was originally based on the<br />

thought that the elephant's trunk evolved to enable the<br />

negatively buoyant animal to breath air from beneath the<br />

surface of the water. The longer the proboscis, the deeper<br />

the animal could forage - eventually resulting in the long,<br />

snorkel-like trunk we see today. This hypothesis has<br />

recently been supported by a study on elephant embryos,<br />

which indicates that the shared ancestor between sirenians<br />

and elephants was an aquatic mammal (Gaeth et al. 1999).<br />

All marine mammals have had to solve the problem<br />

of breathing in an aquatic environment, but the sirenians<br />

have additional unique physical problems: (1) the extant<br />

species are NOT well adapted to cold water; and (2)<br />

sirenians are the only marine mammal herbivores. Only<br />

one extinct species of sirenian was found in extreme cold<br />

waters: Hydrodamalis gigas, commonly known as Steller's<br />

giant sea cow. These animals were three times as large as<br />

manatees, ranging from 25 to 35 feet in length and<br />

weighing up to 8000 pounds. Their extremely large size<br />

enabled them to survive in the frigid Bering Sea where they<br />

feed on giant kelp. All extant species of sirenians are<br />

limited to tropical and sub-tropical waters year round; some<br />

individuals migrate into temperate areas during the<br />

summer. As herbivores, all sirenians are limited to<br />

shallow coastal waters, estuaries, and rivers where aquatic<br />

vegetation is abundant. The West Indian manatee inhabits<br />

riverine and marine systems from Florida to Brazil. But<br />

their distribution is patchy, and appears to be a function of<br />

physical problem solving such as thermoregulation,<br />

foraging, predator avoidance (Hartman 1979), and<br />

osmoregulation. These are the problems we will focus on<br />

here.<br />

Thermoregulation: Water temperature is well<br />

known to be a controlling factor in Florida manatee<br />

distribution in the United States (Hartman 1979, Irvin<br />

1983). Rarely, manatees have been sighted in North<br />

Carolina (Schwartz 1995); they are routinely sighted in<br />

South Carolina and Georgia; but the Chesapeake Bay has<br />

always been considered north of any expected range - even


in the summer. Chessie, a male Florida Manatee, earned<br />

his name by showing up in the Chesapeake Bay during the<br />

summer of 1994. <strong>Sirenian</strong> biologists agreed that Chessie<br />

would die from the cold of oncoming winter if he remained<br />

in the Bay, so they flew him back to Florida, put a satellite<br />

transmitter tag on him and let him go. When the water<br />

began to warm the next summer, Chessie headed north<br />

again. Not only did he temporarily enter the Chesapeake<br />

Bay, but when he came out, he continued his northern<br />

journey up the Atlantic seaboard. Upon reaching Port<br />

Judith, Rhode Island, he finally reversed direction and<br />

began working his way back to Florida for the winter!<br />

[NOTE: For more details go to<br />

www.sirenian.org/chessie.html]<br />

Each winter, hundreds of West Indian manatees<br />

aggregate in Crystal River, Florida, where they are soon<br />

joined by thousands of humans who want to swim with<br />

them. Why does the otherwise elusive manatee tolerate this<br />

human behavior? Why do they keep coming back, year<br />

after year?<br />

These are excellent examples of how manatees have<br />

adapted behaviors to solve the physical problem of<br />

thermoregulation. While most marine mammals have<br />

adapted to cold water by evolving high metabolic rates, the<br />

West Indian manatees have exceptionally low metabolic<br />

rates (Irvin 1983). When metabolic rates are graphed<br />

against body size, most mammals fall along a predictable<br />

curve where the rate decreases as the size increases. If we<br />

compare where manatees should fall on this curve to where<br />

they actually plot out on the graph, we find that manatee<br />

metabolism is only about 20% of what we would expect.<br />

The same comparison with other marine mammals shows<br />

that their metabolic rates are almost twice what we would<br />

expect - enabling them to easily live in cold water. This<br />

physiological problem should limit the West Indian<br />

manatee to warm tropical waters.<br />

In the United States, however, the Florida<br />

subspecies thermoregulates behaviorally by migrating to<br />

both natural and artificial warm water effluents during the<br />

winter months, enabling them to extend their habitat range.<br />

From a proximate cause perspective, the migration<br />

behavior is triggered when the water temperature drops<br />

below 20 degrees Celsius (Hartman 1979; Irvin 1983). But,<br />

from a proximate development perspective, how do adult<br />

manatees know where to find warm water effluents? We<br />

touched on this earlier, when we talked about why manatee<br />

calves remain dependent on their mothers for up to 2 years.<br />

Long term studies of radio tagged females with calves<br />

indicate that manatees initially learn migration routes from<br />

their mothers during the extended parental care period (R.<br />

Bonde, personal communication). Because of this learned<br />

behavior, interruption of man-made thermal effluents may<br />

have negative impacts on manatee survival in areas where<br />

no natural warm water effluents are nearby (Packard et al.<br />

1989).<br />

What about manatees in tropical habitats where<br />

water temperatures are relatively constant? If<br />

7<br />

thermoregulation were the only function of long distance<br />

travel, we would not expect Antillean manatees to exhibit<br />

the same degree of seasonal migration as Florida manatees.<br />

Buddy Powell (personal communication 2000) is working<br />

on that ultimate evolution question through telemetry<br />

studies of Antillean manatees in Belize (see Satellite<br />

Tracking of W.I. Manatees in Belize<br />

www.wesave.org/manatee/). After almost two years of<br />

data, it appears that several manatee mother-calf pairs<br />

remain in the Southern Lagoon area year round. On the<br />

other hand, Greg Smith (personal communication) finds<br />

male manatees seasonally absent on the reef at Basil Jones,<br />

Ambergris Caye, during the winter months of December -<br />

February. In my limited personal observations, I have not<br />

observed manatees on the reef at Gallows Point Reef, near<br />

Belize City, from November through April; but they were<br />

there daily in July and August 1999. Temperature changes<br />

between summer and winter on Gallows Reef vary from<br />

about 28 <strong>–</strong> 32 degrees Celsius (unpublished data), well<br />

above the 20 degree trigger found in Florida manatees. If<br />

temperature is not driving this apparent seasonal migration,<br />

what other factors could be the cause? Greg Smith<br />

hypothesizes that it is driven by seasonal reproductive<br />

cycles. Perhaps the males "hang out" at the reef during the<br />

summer months looking for an estrous female (see<br />

previous section on reproductive problem solving). More<br />

data are required before we can answer these questions.<br />

Foraging: As we explored the bogues (channels)<br />

that snake their way through the mangrove islands off the<br />

coast of Belize, it became apparent that Antillean manatees<br />

preferred certain micro-habitats within the larger habitat<br />

we call the Drowned Cayes. We reliably found manatees<br />

just west of the cayes feeding on turtle grass beds; and we<br />

always found manatees resting in narrow bogues or quiet<br />

coves among the cayes. They tended to travel in the deeper<br />

channels when moving between areas. One of the<br />

parameters of quality manatee habitat is the close<br />

availability of food (Hartman 1979). It appears that<br />

Antillean manatees using the Drowned Cayes are more<br />

likely to rest in areas that are linked to turtle grass beds by<br />

deep channels. Like all sirenians, manatees are<br />

opportunistic herbivores, feeding on a variety of fresh and<br />

saltwater vegetation. Although they may consume fish in<br />

some areas (Powell 1978) and incidentally ingest<br />

invertebrates (Powell 1978; Powell 1984; personal<br />

observations), the main component of their diet is aquatic<br />

vegetation: sea grasses in the marine and estuarine<br />

environment; floating, submerged, and emergent plants in<br />

the riverine environment. Sea grasses, like all plants,<br />

require sunlight for growth, which limits their presence to<br />

relatively shallow coastal waters. From an ultimate<br />

function perspective, the relationship between sea grasses<br />

and water depth has probably prevented manatees from<br />

dispersing into deeper oceanic waters.<br />

I snorkeled up to an Antillean manatee feeding<br />

underwater just west of Swallow Caye <strong>–</strong> at first, I thought it<br />

was dead... I can still recall the adrenalin rush as options


flashed through my mind regarding what to do with a dead<br />

manatee! It was lying perfectly still on the bottom in about<br />

3 meters of water and appeared to be missing its head. As I<br />

floated closer, (heart racing) I began to hear chewing<br />

noises and realized that the manatee had buried its head<br />

into the muddy substrate and was feeding on the sea grass<br />

roots. I soon learned that this is typical of how manatees<br />

feed on the sea grass beds near the Drowned Cayes - eating<br />

both roots and leaves and probably other benthic organisms<br />

living in the mud. Looking for muddy disturbances became<br />

another method of finding the elusive manatee.<br />

If you’ve ever tried to dive down and recover a lost<br />

item in deep water, you know that you, like most mammals,<br />

are positively buoyant and must work to get and stay<br />

submerged. Manatee bones are pachyostoic -- very dense<br />

and lacking marrow -- except in the vertebrae and sternum<br />

(Odell and Reynolds 1991). Because of this, manatees are<br />

negatively buoyant and can lie on the sea bottom without<br />

exerting any energy to stay down. The less energy they<br />

use, the longer manatees can remain submerged between<br />

breaths - making feeding more efficient. Indeed, we think<br />

manatees have the ability to control the volume of air their<br />

lungs, enabling them to rise to the surface, take a breath,<br />

and return to the bottom with no noticeable effort.<br />

Manatees exhibit different problem-solving<br />

behaviors related to foraging in different habitats. In rivers,<br />

manatees are often observed feeding on floating vegetation.<br />

They use their forelimbs like we use our hands to<br />

manipulate aquatic plants towards the mouth. The large<br />

prehensile upper lip is then used to work the plants into the<br />

mouth. When I was observing Georgia and Peaches in<br />

Florida, I was fascinated to see Georgia reach her head out<br />

of the water to feed on plants growing along the shoreline<br />

of the Hontoon Dead River. At times it looked as if she<br />

was going to climb out onto the bank! In fact, Florida<br />

manatees feeding on grasses have been sighted with up to<br />

one third of their body awash (Powell 1984).<br />

Although they are considered opportunistic feeders,<br />

manatees are known to prefer certain species of plants to<br />

others (Bengtson 1981). Hunters of West African manatees<br />

use cassava to lure the animals into box traps (O’Shea<br />

1994). Other traditional hunters attract manatees by<br />

dangling a favorite flower over the water's edge to entice<br />

their approach. While West Indian manatees in Florida and<br />

Belize feed on a variety of vegetation including floating,<br />

submergent, emergent, and over-hanging vegetation, the<br />

Amazonian and the West African manatees feed primarily<br />

on surface vegetation.<br />

In Gambia, when feeding on overhanging<br />

vegetation, West African manatees grasp the leaves of<br />

branches with their lip pads near the water's surface (Powell<br />

1984). They pull the branch into the water; use their<br />

forelimbs to hold it down; and then eat the leaves - but not<br />

the woody petiole or branches. They appear to prefer the<br />

young leaves and shoots of mangroves, as they are known<br />

to return to areas where they have previously fed to crop<br />

any new growth. Amazonian manatees face an even more<br />

8<br />

challenging problem. During the rainy season, floodwaters<br />

allow manatees to literally "forage among the tree tops", but<br />

in the dry season, they are often confined to isolated lakes<br />

and pools devoid of any vegetation. Robin Best calculated<br />

that Amazonian manatees could fast for up to seven months<br />

a year - surviving on stored fat reserves they build up<br />

during the rainy season (O'Shea 1994).<br />

Daryl Domning of Howard University describes<br />

morphological variation in the rostrum deflection among<br />

manatee species and subspecies and suggests that the<br />

variation results from differential foraging behavior among<br />

sirenian species (1980). For example, the rostral deflection<br />

in both the Amazonian and the West African species is<br />

relatively less than it is in the West Indian manatee. These<br />

comparisons of foraging behavior and rostral deflection<br />

represent ultimate perspectives. Comparing foraging<br />

behaviors among the species is an example of ultimate<br />

evolution. Hypothesizing about why the rostral deflection<br />

varies among species (because of different foraging<br />

behaviors) is an example of ultimate function. From an<br />

ultimate perspective, it is interesting to compare the<br />

foraging behavior of manatees to the dugongs. Dugongs<br />

forage exclusively on submerged marine sea grasses and<br />

their rostra are significantly deflected downward when<br />

compared to manatees. The divergence in foraging<br />

behavior between the two families of Sirenia may be due to<br />

the evolution of true grasses in the Caribbean and<br />

continuously regenerating teeth in manatees. Unique to the<br />

Trichechus genus, is the fact that manatees generate new<br />

molars throughout their lifespan (Domning and Hayek<br />

1984). As older molars are worn down from the abrasive<br />

silica content in true grasses, new molars gradually move<br />

forward at the rate of a few millimeters per month. The<br />

forward most molars eventually fall out and are replaced<br />

from the rear in a horizontal manner. This would be<br />

analogous to humans continuously generating new wisdom<br />

teeth at the rate of four - 2 lower and 2 upper - every few<br />

months! Daryl Domning (1982) convincingly argues that<br />

this trait enabled manatees to out compete dugongs in the<br />

Atlantic a few million years ago. While manatees are only<br />

found in the Atlantic and continue to forage on a variety of<br />

vegetation, dugongs are only found in the Indo-Pacific and<br />

forage exclusively on marine sea grasses.<br />

Predation: Although they may have existed in the past, we<br />

know of few natural predators on modern West Indian<br />

manatees...EXCEPT for humans. Large aquatic predators<br />

(crocodiles, alligators, sharks, and hippopotamus) have<br />

been hypothesized to take the occasional small or weak<br />

animal (Odell 1982; Powell 1984). But, in one of the few<br />

documented cases, Johnson (1937, as reported in Powell<br />

1984) found that only one out of one hundred crocodiles cut<br />

open contained the remains of a manatee.<br />

Amazonian manatees, on the other hand, still have to<br />

contend with predation by aquatic and terrestrial carnivores<br />

such as jaguars, caimans, and sharks (Reynolds and Odell<br />

1991) <strong>–</strong> especially during the dry season when then are<br />

stranded by receding flood waters. Both fossil and historical


ecords indicate that manatees have been hunted both for<br />

subsistence and commercially throughout the history of<br />

humans (Lefebvre et al. 1989, Reynolds and Odell 1991).<br />

While illegal poaching still exists <strong>–</strong> especially in remote<br />

areas -- most modern predation is incidental <strong>–</strong> resulting<br />

from entanglement in fishing gear, shark nets, and water<br />

control devices, and from collisions with watercraft. From<br />

an ultimate function perspective, we may hypothesize that<br />

the reason manatees are elusive creatures is to avoid<br />

predation. In other words, those animals that inherited a<br />

natural tendency towards elusive behavior were more<br />

reproductively successful.<br />

The story of Steller’s sea cow demonstrates how<br />

quickly humans can extirpate a species, particularly when<br />

population numbers are already reduced. The sea cow was<br />

discovered and described by Georg Wilhelm Steller, a<br />

German naturalist assigned to Captain Vitus Bering during<br />

a Russian Expedition to Alaska (Steller 1988). This giant<br />

sirenian was 25-35 feet long and weighed ~8000 pounds<br />

with flukes than spanned 8 feet. Unlike modern sirenians,<br />

it lived in extremely cold waters in the North Pacific. It had<br />

no teeth, but two grooved plates - one upper and one lower<br />

with which it crunched giant sea kelp. During the summer<br />

of <strong>17</strong>41, Captain Bering set sail from Kamchatka in NE<br />

Russia with 2 ships, the St. Peter and the St. Paul. During<br />

the voyage, a storm separated his ships; Bering, Steller, and<br />

the crew of the St. Peter were shipwrecked on an unknown<br />

island (later named Bering Island). Although Bering did<br />

not survive the winter, Steller and many of the crew did. In<br />

his book, A Voyage with Bering, Steller credits their<br />

survival to the giant sea cow. Only after the crew learned<br />

to hunt the sea cow did they begin to regain the strength to<br />

repair their ship. When they returned to Kamchatka in the<br />

summer of <strong>17</strong>42, they told of the wonderful sea cow meat.<br />

New hunting expeditions were formed almost immediately<br />

and every year thereafter. The expeditions would return to<br />

Bering Island where they spent 8-9 months hunting furanimals<br />

and eating sea cow meat to survive. Indeed, many<br />

of the expeditions are reported to have wintered on Bering<br />

Island for the express purpose of collecting sea cow meat to<br />

provision their ships for the rest of their 3-4 year voyage to<br />

America. As a result, the last sea cow was reported killed<br />

in <strong>17</strong>68, only 27 years after modern humans had discovered<br />

the island and the species.<br />

Even before Russian sailors were hunting Steller’s<br />

sea cow to extinction, European buccaneers and explorers<br />

were provisioning their ships with Antillean manatee meat,<br />

which they harvested themselves or purchased from the<br />

indigenous people (Reynolds and Odell 1991). In the<br />

Panama area, it is estimated that at one time, 7-8 thousand<br />

manatees were being harvested annually. The Amazonian<br />

manatee continued to be commercially harvested for it's<br />

tough skin (which was made into leather products) into the<br />

1950's. From an ultimate function perspective, it stands to<br />

reason that manatees that behaved in an elusive manner<br />

lived to produce more offspring during the last several<br />

centuries.<br />

9<br />

The greatest documented predation on manatees<br />

today occurs incidentally in Florida due to competition for<br />

space between manatees and humans. Government<br />

agencies have been documenting manatee mortality in<br />

Florida for almost three decades. The proportion of deaths<br />

related to collisions with watercraft tends to increase each<br />

year as more and more people move to Florida. For current<br />

statistics on Florida manatee mortality, visit the FMRI and<br />

FFWCC websites at www.fmri.usf.edu/manatees.htm and<br />

www.state.fl.us/fwc/psm/manatee/manatee.htm. Even the<br />

most elusive manatee has a difficult time avoiding<br />

interaction with people as the human population and<br />

development continue to grow.<br />

Osmoregulation: Unlike the Amazonian manatee,<br />

which is endemic to the fresh waters of the Amazon River<br />

basin, and the dugong, which is only found in marine<br />

habitats, the West Indian and West African manatees<br />

appear to move freely between fresh and marine<br />

environments. Although the Florida subspecies is usually<br />

associated with fresh or brackish water, it is occasionally<br />

found far offshore in high salinity water (Reynolds and<br />

Ferguson 1984). Large amounts of barnacle growth suggest<br />

that some individuals spend prolonged periods in marine<br />

environments (Husar 1977, Hartman 1979). Antillean<br />

manatees are found year round in totally marine<br />

environment of red mangrove islands in Belize (personal<br />

observation). How does this species osmoregulate as it<br />

moves among fresh, brackish, and saltwater? Two<br />

alternative hypotheses come to mind: (1) West Indian<br />

manatees have physiological adaptations that enable them<br />

to maintain water balance and/or (2) W. I. Manatees<br />

behaviorally maintain water balance by seeking out fresh<br />

water sources in marine environments. Graham Worthy, a<br />

physiologist at Texas A&M University, and his students are<br />

working on this and proximate cause questions involving<br />

physiology. Preliminary results indicate that W. I.<br />

Manatees are “good osmoregulators regardless of the<br />

environment” (Ortiz et al. 1998).<br />

A FEW THOUGHTS…<br />

As we examine problem-solving behavior in the<br />

West Indian manatee using the ethological perspectives of<br />

cause, development, evolution, and function, we begin to<br />

realize how many questions remain un-answered about this<br />

elusive marine mammal. Although we know (from the<br />

study of other mammals) that the proximate cause of<br />

behavior is a complex interaction between internal<br />

mechanisms and external stimuli, we don't know the<br />

specific triggers for many manatee behaviors. Proximate<br />

development has not been well studied for several reasons.<br />

Free ranging female manatees with new calves tend to<br />

isolate themselves in secluded areas making behavior<br />

difficult to observe. Observations of captive raised<br />

manatees may not be indicative of normal development.<br />

The West Indian manatee is an endangered species making<br />

experimental manipulation difficult, yet extremely<br />

important to the successful rehabilitation and release of


injured manatees. For example, how will a captive raised<br />

calf learn to find warm water effluents during cold spells?<br />

Can adult manatees learn successful migration routes or<br />

must they be learned during early development? From an<br />

ultimate perspective, manatees are also quite challenging<br />

due to the lack of closely related extant species and to the<br />

sparse fossil record. But, paleo-sirenian research by Daryl<br />

Domning, and others, continues to offer insight to the<br />

evolution and function of modern manatee behaviors.<br />

Why are there fewer sirenian species today than during the<br />

past? Does the evolution of the species Homo correlate<br />

with the decline of sirenians or were other environmental<br />

factors the reason for their extinctions? Will answers to<br />

these and other ultimate questions aid in our conservation<br />

efforts?<br />

Although manatees are generally considered elusive,<br />

there are cases where they appear to be curious and actually<br />

initiate contact with humans. Likewise, many behaviors<br />

tend vary between individuals, populations, and species.<br />

Because of the variable nature of manatee behavior, we<br />

must be careful in applying what we know about the Florida<br />

population to other areas. The Florida subspecies is<br />

fortunate to have the efforts of many US citizens and<br />

agencies working toward conservation issues and manatee<br />

behavior plays and important role in making management<br />

decisions in Florida. However, West Indian manatees are<br />

considered endangered throughout their range. Continued<br />

research effort on populations in the more tropical regions<br />

of the Caribbean is necessary for decision makers in those<br />

countries to make effective management decisions.<br />

REFERENCES<br />

Alcock, John. 1998. Animal Behavior. Sixth edition.<br />

Sunderland, Massachusetts: Sinauer Associates,<br />

Inc. 640 pp.<br />

Anderson, Paul K. 1997. Shark Bay dugongs in summer. I:<br />

Lek mating. Behaviour. 134(5-6):433-462.<br />

Auil, Nicole. 1998. Belize Manatee Recovery Plan.<br />

UNDP/GEF Coastal Zone Management Project,<br />

BZE/92/G31, Belize/UNEP Caribbean<br />

Environment Programme, Kingston, Jamaica. 67<br />

pp.<br />

Bengtson, John L. 1981. Ecology of Manatees Trichechus<br />

manatus) in the St. Johns River, Florida. Ph. D.<br />

thesis. University of Minnesota.<br />

Bengtson, John L., and Shannon M. Fitzgerald. 1985.<br />

Potential role of vocalizations in West Indian<br />

manatees. Journal of Mammalogy. 66(4):816-<br />

819.<br />

Domning, Daryl P. 1980. Feeding position preference in<br />

manatees (Trichechus). Journal of Mammalogy<br />

61(3):544-547.<br />

Domning, Daryl P. 1982. Evolution of manatees: a<br />

speculative history. Journal of Paleontology.<br />

56(3):599-619.<br />

10<br />

Domning, Daryl P., and Lee-Ann Hayek. 1984. Horizontal<br />

tooth replacement in the Amazonian manatee.<br />

Mammalia 48:105-127.<br />

Drickamer, Lee C., Stephen H. Vessey, and Doug Meikle.<br />

1996. Animal Behavior: Mechanisms, Ecology,<br />

Evolution. Fourth edition. Boston: William C.<br />

Brown Publishers. 447 pp.<br />

Fischer, M. S. 1990. The unique ear of elephants and<br />

manatees (Mammalia): a phylogenetic paradox.<br />

C. R. Acad. Sci., Paris, SER. III, vol. 31, no. 4, pp.<br />

157-162.<br />

Gaeth, A. P., R. V. Short and M. B. Renfree. 1999. The<br />

developing renal, reproductive, and respiratory<br />

systems of the Africa elephant suggest an aquatic<br />

ancestry. Proceedings of the National Academy of<br />

Sciences of the United States of America<br />

96(10):5555-5558.<br />

Hartman, Daniel S. 1979. Ecology and behavior of the<br />

manatee (Trichechus manatus) in Florida. Special<br />

publication no. 5. American Society of<br />

Mammalogists. 153 pp. ISBN 0-9436-1204-7.<br />

Hernandez, Patricia, John E. Reynolds, III, Helene Marsh,<br />

and Miriam Marmontel. 1995. Age and<br />

seasonality in spermatogenesis of Florida<br />

manatees. pp. 84-95 in Thomas J. O'Shea, Bruce<br />

B. Ackerman, and H. Franklin Percival, editors.<br />

Population biology of the Florida manatee.<br />

National Biological Service Information and<br />

Technology Report 1.<br />

Husar, Sandra L. 1977. The West Indian manatee<br />

Trichechus manatus. Wildlife Research Report<br />

No. 7. U. S. Department of the Interior, Fish and<br />

Wildlife Service, Washington, D.C. 22 pp.<br />

Irvin, A. Blair. 1983. Manatee metabolism and its<br />

influence on distribution in Florida. Biological<br />

Conservation. 25:315-334.<br />

Koelsch, Jessica K. 1997. The Seasonal Occurrence and<br />

Ecology of Florida Manatees (Trichechus manatus<br />

latirostris) in Coastal Waters near Sarasota,<br />

Florida. Masters thesis. University of South<br />

Florida. Tampa, Florida.<br />

Lefebvre, Lynn W., Thomas J. O'Shea, Galen B. Rathbun,<br />

and R. C. Best. 1989. Distribution, status, and<br />

biogeography of the West Indian manatee. pp.<br />

567-610 in Biogeography of the West Indies. ed.<br />

Charles A. Woods. Gainesville, FL: Sandhill<br />

Crane Press. ISBN 0-8493-2001-1.<br />

Lehner, Philip N. 1996. Handbook of Ethological<br />

Methods, 2 nd edition. Cambridge University Press.<br />

672 pp.<br />

Maluf, N. S. R. 1995. Kidney of elephants. Anatomical<br />

Record 242(4):491-514.<br />

Marmontel, Miriam. 1995. Age and reproduction in<br />

female Florida manatees. Pages 98-119 in Thomas<br />

J. O'Shea, Bruce B. Ackerman, and H. Franklin<br />

Percival, editors. Population biology of the


Florida manatee. National Biological Service<br />

Information and Technology Report 1.<br />

Marmontel, Miriam, Thomas J. O'Shea, and S.R.<br />

Humphrey. 1990. An evaluation of bone growthlayer<br />

counts as an age determination technique in<br />

Florida manatees. National Technical Information<br />

Service, Springfield, Va. Document PB 91-<br />

103564. 104 pp.<br />

Marsh, Helene. 1995. The life history, pattern of breeding,<br />

and population dynamics of the dugong. pp. 75-<br />

83 in Thomas J. O'Shea, Bruce B. Ackerman, and<br />

H. Franklin Percival, editors. Population<br />

biology of the Florida manatee. National<br />

Biological Service Information and Technology<br />

Report 1.<br />

Martin, Paul, and Patrick Bateson. 1993. Measuring<br />

Behaviour <strong>–</strong> An introductory guide, 2 nd edition.<br />

Cambridge University Press. 222 pp.<br />

Odell, Daniel K. 1982. West Indian Manatee Trichechus<br />

manatus. in Wild Mammals of North America<br />

Biology, Management, and Economics. Joseph A.<br />

Chapman and George A. Feldhamer, editors. The<br />

Johns Hopkins University Press. Baltimore and<br />

London.<br />

O'Shea, Thomas J. 1994. Manatees. Scientific American.<br />

July 1994.<br />

O'Shea, Thomas J. Bruce B. Ackerman, and H. Franklin<br />

Percival. 1995. Population Biology of the<br />

Florida manatee. National Biological Service<br />

Information and Technology Report 1.<br />

O’Shea, Thomas J., and Wayne Hartley. 1995.<br />

Reproduction and early-age survival of manatees<br />

at Blue Spring, upper St. Johns River, Florida. pp.<br />

157-<strong>17</strong>0 in Thomas J. O'Shea, Bruce B.<br />

Ackerman, and H. Franklin Percival, editors.<br />

Population biology of the Florida manatee.<br />

National Biological Service Information and<br />

Technology Report 1.<br />

O'Shea, Thomas J. and Charles A. 'Lex' Salisbury. 1991.<br />

Belize - a last stronghold for manatees in the<br />

Caribbean. Oryx. 25(3): 156-164.<br />

Ortiz, Rudy M., Graham A. J. Worthy and Duncan S.<br />

MacKenzie. 1998. Osmoregulation in wild and<br />

captive West Indian manatees (Trichechus<br />

manatus). Physiological Zoology 71(4):449.457.<br />

Packard, Jane M., R. Kipp Frohlich, John E. Reynolds, III,<br />

and J. Ross Wilcox. 1989. Manatee response to<br />

interruption of a thermal effluent. Journal of<br />

Wildlife Management 53(3):692-700.<br />

Powell, James A. 1978. Evidence of carnivory in manatees<br />

(Trichechus manatus). Journal of Mammalogy.<br />

59(2):442.<br />

Powell, James A. 1984. Manatees in the Gambia River<br />

Basin and potential impact of the Balingho<br />

Antisalt Dam with notes on Ivory Coast, West<br />

Africa. Trip Report. Institute for Marine Studies.<br />

11<br />

University of Washington. Seattle Washington<br />

98195.<br />

Ralph, C. L., S. Young, R. Gettinger and T. J. O’Shea.<br />

1985. Does the manatee have a pineal body?<br />

Acta Zoologica 66(1):55-60.<br />

Rathbun, Galen B., James P. Reid, Robert K. Bonde, and<br />

James A. Powell. 1995. Reproduction in freeranging<br />

Florida manatees. pp. 135-156 in T.J.<br />

O'Shea, B.B. Ackerman, and H.F. Percival,<br />

editors. Population biology of the Florida<br />

manatee. National Biological Service Information<br />

and Technology Report 1.<br />

Reynolds, John E., III. 1981. Aspects of the social<br />

behaviour and herd structure of a semi-isolated<br />

colony of West Indian manatees, Trichechus<br />

manatus. Mammalia. 45(4):431-451.<br />

Reynolds, John E., III, and John C. Ferguson. 1984.<br />

Implications of the presence of manatees<br />

(Trichechus manatus) near the Dry Tortugas<br />

islands. Florida Scientist 47(3):187-189.<br />

Reynolds, John E. III, and Daniel K. Odell. 1991.<br />

Manatees and Dugongs. New York: Facts on File,<br />

Inc. 192 pp. ISBN 0-8160-2436-7.<br />

Schwartz, F. J. 1995. Florida manatees, Trichechus<br />

manatus (Sirenia: Trichechidae) in North Carolina<br />

1919-1994. Brimleyana 22:53-60.<br />

Springer, M. S., G. C. Cleven, O. Madsen, W. W. De Jong,<br />

V. G. Waddell, H. M. Amrine, M. J. Stanhope.<br />

1997. Endemic African mammals shake the<br />

phylogenetic tree. Nature (London) 388<br />

(6637):61-64.<br />

Steller, Georg Wilhem. 1988. Journal of a Voyage with<br />

Bering <strong>17</strong>41-<strong>17</strong>42. Stanford, California: Stanford<br />

University Press. 252 pp.<br />

Stejneger, Leonhard. 1887. How the great northern seacow<br />

(Rytina) became exterminated. The American<br />

Naturalist 21(12):1047-1054.<br />

RECOMMENDED READING<br />

Caldwell, David K., and Melba C. Caldwell. 1985.<br />

Manatees - Trichechus manatus, Trichechus<br />

senegalensis, and Trichechus inunguis. pp. 33-36<br />

in Handbook of Marine Mammals, vol. 3, The<br />

<strong>Sirenian</strong>s and Baleen Whales. eds. S. H.<br />

Ridgway and R. J. Harrison. London: Academic<br />

Press. ISBN 0-1258-8503-2.<br />

Hartman, Daniel S. 1979. Ecology and behavior of the<br />

manatee (Trichechus manatus) in Florida. Special<br />

Publication no. 5. American Society of<br />

Mammalogists. 153 pp. ISBN 0-9436-1204-7.<br />

Reynolds, John E. III, and Daniel K. Odell. 1991.<br />

Manatees and Dugongs. New York: Facts on File,<br />

Inc. 192 pp. ISBN 0-8160-2436-7.


Zeiller, Warren. 1992. Introducing the Manatee.<br />

Gainesville, Florida: University Press of Florida.<br />

151 pp. ISBN 0-8130-1152-3.<br />

World Wide Web Internet URLs:<br />

Animal Behavior by Dr. Jane M. Packard:<br />

www.tamu.edu/ethology/<br />

Call of the Siren by Caryn Self Sullivan:<br />

www.sirenian.org/caryn.html<br />

The Nobel Prize in Physiology or Medicine 1973 <strong>–</strong> Karl<br />

von Frisch, Konrad Lorenz, Nikolaas Tinbergen, for their<br />

discoveries concerning organization and elicitation of<br />

individual and social behaviour patterns:<br />

www.nobel.se/medicine/laureates/1973/index.html<br />

Satellite Tracking West Indian Manatees in Belize:<br />

www.wesave.org/manatee/<br />

Florida Fish and Wildlife Conservation Commission:<br />

www.state.fl.us/fwc/psm/manatee/manatee.htm<br />

Florida Marine Research Institute:<br />

www.fmri.usf.edu/manatees.htm<br />

USGS Sirenia Project:<br />

www.fcsc.usgs.gov/sirenia/<br />

Sea World Education Department:<br />

www.seaworld.org/manatee/manatees.html<br />

Florida Power & Light Company:<br />

www.dep.state.fl.us/psm/webpages/manatees/booklet.html<br />

University of Wisconsin - The Brain of the Florida<br />

Manatee:<br />

www.neurophys.wisc.edu/manatee/<br />

University of Florida - Manatee Research Group:<br />

www.vetmed.ufl.edu/ufmrg/manatee/<br />

GLOSSARY<br />

agonistic: a general term that includes aggressive,<br />

submissive, and defensive behaviors that appear when the<br />

adrenal hormones are activated.<br />

anecdote: a short story or observation which has not been<br />

"tested" by the scientific method.<br />

anthropocentric: considering human beings as the most<br />

significant entity of the universe; interpreting or regarding<br />

the world in terms of human values and experiences.<br />

anthropogenic: of, relating to, or resulting from the<br />

influence of human beings on nature.<br />

anthropomorphic: described or thought of as having a<br />

human form or human attributes; ascribing human<br />

characteristics to nonhuman things.<br />

bogue: the local term for a channel of water that flows<br />

through a mangrove caye in Belize, C.A.<br />

cause: stimulus outside the animal plus the internal<br />

physiological state of the animal.<br />

Cetacea: the order of aquatic mostly marine mammals that<br />

include the whales, dolphins, porpoises, and related forms<br />

and that have a torpedo-shaped nearly hairless body,<br />

paddle-shaped forelimbs but no hind limbs, one or two<br />

nares opening externally at the top of the head, and a<br />

horizontally flattened tail used for locomotion.<br />

convergent: similar traits in species from very different<br />

genetic lineages, due to similar environmental functions.<br />

12<br />

development: changes in behavioral traits as an individual<br />

ages.<br />

divergent: different traits in species from similar genetic<br />

lineages due to differences in their environments.<br />

endemic: restricted or peculiar to a locality or region.<br />

Amazonian manatees are only found in the Amazon River<br />

Basin.<br />

ephemeral: short period of existence; opposite of eternal.<br />

estrous: the period during which a female has produced an<br />

egg ready to fertilize and is receptive to copulation.<br />

evolution: change in proportion of genotypes within the<br />

gene pool of a population over many many generations.<br />

extant: currently or actually existing; not destroyed or lost.<br />

extinct: no longer existing.<br />

extirpated: to destroy completely.<br />

folk psychology: a non-scientific way of talking about<br />

animal behavior. Uses human terms to describe non-human<br />

animal behavior. Intuitive <strong>–</strong> based on anecdotes,<br />

experiences, observations. Anthropomorphic. Assumes<br />

that animals have beliefs, feelings, desires.<br />

function: the meaning of a behavior in terms of survival<br />

and reproduction in a given environment.<br />

gestation: the carrying of young in the uterus.<br />

herbivores: animals that feed exclusively on plants.<br />

individual: a single organism; a way of understanding the<br />

biological hierarchy of concepts - those pertaining to<br />

physiological processes and experiences of each organism;<br />

in contrast to population, which focuses more on gene<br />

pools of groups of animals that interbreed.<br />

lactation: the secretion of milk.<br />

manatee hole: concave depression in sandy or muddy<br />

substrate where manatees habitually rest.<br />

mating herd: the name given an estrous female and the<br />

aggregation of males which follow her around attempting to<br />

mate.<br />

mating system: males strategies and female strategies as<br />

observed in a population.<br />

morphology: the form and structure of an organism or any<br />

of its parts.<br />

natural selection: the process that produces evolutionary<br />

changes in a population due to heritable traits in some<br />

individuals which result in those individuals being more<br />

reproductively successful; the process only occurs IF (1)<br />

there are variances in traits within a population, (2) the<br />

variances are heritable, (3) individuals with certain<br />

variances have greater reproductive success than<br />

individuals with other variances, then over time, we would<br />

expect to see the trait selected "for" become more common<br />

in the population.<br />

neonate: term used distinguish a newborn cetacean or<br />

sirenian calf from an older calf <strong>–</strong> you can still see the fetal<br />

folds on a neonate and the skin is usually darker than on an<br />

older calf.<br />

osmoregulation: regulation of osmotic pressure especially<br />

in the body of a living organism. how an organism<br />

regulates the amount of water entering and leaving its body.


parturition: the action or process of giving birth to<br />

offspring; birth, whelping.<br />

pinniped: any of a suborder (Pinnipedia) of aquatic<br />

carnivorous mammals (as a seal or walrus) with all four<br />

limbs modified into flippers.<br />

polyandry: a mating system where one female mates with<br />

two or more males at a time. (adj: polyandrous)<br />

population: a group of animals that interact and interbreed.<br />

postpardum: following parturition; after giving birth.<br />

prehensile: an appendage adapted for seizing or grasping<br />

especially by wrapping around - examples: a monkey has a<br />

prehensile tail and an elephant has a prehensile trunk.<br />

proboscis: the trunk of an elephant; also any long flexible<br />

snout; any of various elongated processes of the oral region<br />

of an invertebrate.<br />

promiscuity: a mating system where there is no prolonged<br />

association between the mating pair and at least one sex<br />

engages in multiple mates. (adj: promiscuous)<br />

proximate: immediately preceding or following as in a<br />

chain of events, causes, or effects; in ethology - the<br />

perspective that looks at the cause and development of<br />

behavior at the individual animal level.<br />

13<br />

scramble-polygyny: a mating system where many males<br />

try to mate with many females, but without overt<br />

competition among themselves.<br />

Sirenia: the taxonomic order of aquatic herbivorous<br />

mammals including the manatee, dugong, and Steller's sea<br />

cow. <strong>Sirenian</strong>s are members of the Order Sirenia.<br />

social animals: tending to form cooperative and<br />

interdependent relationships with others of one's kind;<br />

living and breeding in more or less organized communities.<br />

taxonomy: orderly classification of plants and animals<br />

according to their presumed natural relationships; the study<br />

of the general principles of scientific classification;<br />

systematics.<br />

thermoregulate: the maintenance or regulation of<br />

temperature; specifically the maintenance of a particular<br />

temperature of the living body.<br />

ultimate: most remote in space or time; last in a<br />

progression or series; in ethology - the perspectives of<br />

evolution and function that look at behavior at the<br />

population and/or species level.<br />

uniparous: producing one offspring at a time.<br />

ACKNOWLEDGMENTS<br />

Financial support for my research on manatees comes from the National Science Foundation Graduate Fellowship<br />

Program (1998-2001), the Oceanic Society of San Francisco (1998-1999), the Lerner-Gray Fund for Marine Research at the<br />

American Museum of Natural History (2000), and Earthwatch Institute (2001). A special thank you goes out to all the<br />

volunteers who have participated in field research with me in Belize. Many sirenian researchers from around the world have<br />

been quite generous in answering the multitude of questions I asked of them. Special thanks go to Jane Packard and Bill<br />

Evans of Texas A&M University, Bob Bonde of the USGS Sirenia Project, and Daryl Domning of Howard University, for<br />

their perpetual availability via email. I thank Nicole Auil of the Belize Coastal Zone Management Institute and Authority;<br />

Sidney Turton, Elda Cabellos, and the staff of Spanish Bay Resort; Greg Smith of the Belize National Manatee Working<br />

Group (BNMWG); and Buddy Powell of the Florida Marine Research Institute & the BNMWG for their continuing support.<br />

Most importantly, I offer an extra special thank you to my primary boat operator, Armando "Patch" Muñoz, for his hours of<br />

tolerance and patience as we searched for the Elusive Manatee among the Drowned Cayes in Belize. Many others have<br />

helped me along the way by taking me into the field or assisting me in the field, sharing their personal knowledge and<br />

experiences, donating slides to my educational programs, directing me towards the proper references, and just by being there<br />

to listen. In no particular order, they include: Maxine Monsanto, Heidi Petersen, Kate Schafer, Eleno “Landy” Requena,<br />

Kecia Kerr, Barbara Bilgre, Katie LaCommare, Leszek Karcmarski, Birgit Winning, Jaime Gilardi, Beth Wright, Lenisa<br />

Tipton, Tami* Gilbertson, Jessica Koelsch, Graham Worthy, Angela Garcia-Rodriguez, Mike Bragg, Edmund Gerstein, Rich<br />

Harris, Lizz Singh, Chocolate Heredia, Hans Rothauscher, Joe Olson, Patti Thompson, Bruce Ackerman, Cathy Beck, Chip<br />

Deutsch, Lynn Lefebvre, Tom O'Shea, Tom Pitchford, Roger Reep, Wally Welker, John Reynolds, Butch Rommel, Peter<br />

Tyack, Andrea Gill, Kevin Andrewin, Elaine Perez, Rob Young, and my extended family of relatives and friends.<br />

Want to learn more about manatees? Visit my homepage: www.sirenian.org/caryn.html


Aquatic Mammals 2003, 29.3, 342<strong>–</strong>354<br />

Seasonal occurrence of male Antillean manatees (Trichechus manatus<br />

manatus) on the Belize Barrier Reef<br />

Caryn Self-Sullivan 1,2 , Gregory W. Smith 3 , Jane M. Packard 1,2 and<br />

Katherine S. LaCommare 2,3<br />

1 Texas A&M University, Department of Wildlife & Fisheries Sciences, Mail Stop 2258, College Station,<br />

Texas 77843<strong>–</strong>2258, USA<br />

2 <strong>Sirenian</strong> <strong>International</strong>, Inc., 200 Stonewall Drive, Fredericksburg, Virginia 22401<strong>–</strong>2110, USA<br />

3 P.O. Box 142, San Pedro Town, Belize, Central America<br />

4 University of Massachusetts-Boston, Department of Biology, Boston, Massachusetts 02125, USA<br />

Abstract<br />

A fragment of manatee habitat that crosses the<br />

border of Belize and Mexico includes both activity<br />

centres and travel routes linking rivers, lagoons,<br />

seagrass beds and mangrove islands near Chetumal<br />

Bay. Little is known about how geophysical<br />

features like coral reefs may in uence manatee<br />

movements within and between habitat fragments<br />

like this. In this inductive study (1995<strong>–</strong>2001), we<br />

documented the seasonal occurrence of Antillean<br />

manatees at breaks in the northern Belize Barrier<br />

Reef. Survey locations were at: (1) Bacalar Chico<br />

National Park and Marine Reserve on Ambergris<br />

Caye (Basil Jones Cut) and (2) breaks in the reef<br />

70 km south near the Drowned Cayes (Gallows<br />

Reef). The probability of sighting at least one<br />

manatee on a 20-min point scan survey was 40%<br />

(n=382). Sighting probability was signi cantly<br />

higher during the summer season (May<strong>–</strong>August)<br />

compared to winter months (December<strong>–</strong>March).<br />

Group size ranged from one to ve manatees,<br />

peaking earlier (May) at the northern than southern<br />

site (August). Seventeen identi able individuals<br />

accounted for 87% of the sightings at Basil Jones<br />

Cut, with re-sightings within and between years.<br />

One individual from Basil Jones Cut was re-sighted<br />

at Gallows Reef. Of the manatees for which sex was<br />

determined, 100% were males. No calves were<br />

sighted. To better understand manatee activity<br />

centres and travel routes, we identi ed potential<br />

hypotheses relating seasonal in uences, stopover<br />

sites for travelling males, and habitat connectivity.<br />

To protect this highly vulnerable species, we recommend<br />

inclusion of the Belize Barrier Reef as an<br />

important component of manatee habitat within<br />

the coastal zone of Belize.<br />

Key words: Antillean manatee, Trichechus manatus<br />

manatus, Caribbean, Belize, habitat connectivity,<br />

? 2003 EAAM<br />

stopover sites, coastal zone management, Belize<br />

Barrier Reef, fragmented populations, seasonal<br />

habitat use.<br />

Introduction<br />

A sub-species of the West Indian manatee, the<br />

Antillean manatee (Trichechus manatus manatus)<br />

occurs in rivers and coastal marine systems of at<br />

least 19 countries in the Wider Caribbean Region,<br />

including the Greater Antilles, Mexico, Central<br />

America, and South America (CEP/UNEP, 1995;<br />

Lefebvre et al., 2001). It is listed by the IUCN<br />

(Hilton-Taylor, 2001) as vulnerable, in continuing<br />

decline, with severely fragmented populations (VU<br />

A1cd, C2a), and has been identi ed as one of the<br />

‘priority protected species of regional concern’<br />

(CEP/UNEP, 1995 pp. 1).<br />

One population of this focal species spans the<br />

border of Belize and Mexico, a diverse habitat<br />

including Chetumal Bay and the northern Belize<br />

Barrier Reef Lagoon System (Fig. 1). With a<br />

relatively short coastline extending from the Gulf<br />

of Honduras in the south to Chetumal Bay in<br />

the north, Belize reports the largest number of<br />

Antillean manatees in the Caribbean region<br />

(O’Shea & Salisbury, 1991). The contiguous<br />

Chetumal Bay is one of the most important areas<br />

for manatees in Mexico (Morales et al., 2000) and<br />

Northern Belize (Auil, 1998). Primary habitat consists<br />

of rivers, coastal lagoons and bays, and mangrove<br />

islands between the Belize Barrier Reef and<br />

the mainland (Bengtson & Magor, 1979; O’Shea &<br />

Salisbury 1991; Gibson, 1995; Auil, 1998; Morales-<br />

Vela et al., 2000). OVshore atoll systems, such as<br />

TurneVe Atoll, are considered secondary habitat for<br />

manatees (Gibson, 1995; Auil, 1998; Morales-Vela<br />

et al., 2000).


Manatees on the Belize Barrier Reef<br />

Figure 1. Northern Belize Barrier Reef Lagoon System and Southern Chetumal Bay showing survey<br />

locations and the 100-fathom line (100 fathoms=183 m, map modi ed from Purdy et al., 1975).<br />

343


344 C. Self-Sullivan et al.<br />

As identi ed by Packard & Wetterqvist (1986)<br />

in Florida, the components of manatee habitat<br />

systems include activity centres, travel routes,<br />

resources for expansion (potential feeding areas for<br />

recovering populations), essential areas (necessary<br />

to survive seasonal extremes), and the supporting<br />

ecosystem. For the purpose of the present study, we<br />

focused on the former two. We de ned activity<br />

centres as areas where manatees were frequently<br />

observed using resources such as vegetation and<br />

freshwater, in all seasons. For Belize, previous<br />

studies (Gibson, 1995; Lefebvre et al., 2001) indicated<br />

activity centres were located in silty substrates<br />

at 1<strong>–</strong>3 m depth. Activity centres could have<br />

attracted both residents and travellers (Reid et al.,<br />

1991; Koelsch, 1997). We de ned travel routes as<br />

connections between activity centres used by individual<br />

manatees for daily, seasonal, or migratory<br />

movements (Sanderson, 1966). We were interested<br />

in whether geophysical characteristics such as a reef<br />

could have been used as landmarks by traveling<br />

manatees.<br />

Traditional knowledge in local communities indicated<br />

manatees were often sighted on the Belize<br />

Barrier Reef in the summer. However, the reef had<br />

not been included in descriptions of manatee habitat<br />

in Belize, possibly due to a bias from research<br />

done in Florida where reefs were not present in<br />

areas where manatees had been studied. Only one<br />

section of Belize Barrier Reef was included in<br />

national standardized surveys i.e., in the north<br />

where it lies within several 100 m of Ambergris<br />

Caye and has been classi ed as the caye habitat<br />

type (Auil, 1998). As recommended by Weeks &<br />

Packard (1997), we listened to local residents and<br />

chose to document in a systematic manner the<br />

patterns of manatee occurrence that they perceived.<br />

Using an inductive approach, we examined whether<br />

the seasonal trend was robust, and explored the<br />

reasons that this component of the habitat system<br />

would be most attractive to manatees in the<br />

summer.<br />

In this paper, we describe the seasonal occurrence<br />

of predominately male manatees at two locations<br />

along the Belize Barrier Reef: (1) Basil Jones Cut,<br />

which was isolated by Ambergris Caye from a<br />

manatee activity centre in Chetumal Bay, and (2)<br />

Gallows Reef, which was adjacent to a manatee<br />

activity centre in the Belize Barrier Reef lagoon<br />

system (the Drowned Cayes) and exposed to frequent<br />

boat traYc. The study was initiated at Basil<br />

Jones Cut and extended to Gallows Reef. If<br />

manatee presence on the reef was in uenced by<br />

access to warm or fresh water during the winter, we<br />

expected to nd manatees at Gallows Reef when<br />

they were not at Basil Jones. Similarly, if the<br />

predominance of males was in uenced by isolation<br />

from an activity centre, we expected to nd female<br />

manatees and calves at Gallows Reef when they<br />

were not at Basil Jones.<br />

Materials and Methods<br />

Study site<br />

The Belize Barrier Reef extends from the Mexican<br />

border, where it is only a few meters from the coast,<br />

to the Gulf of Honduras where it is 50 km oVshore<br />

(Purdy et al., 1975). It forms an important geophysical<br />

barrier between the shallow coastal lagoon<br />

system and the deep Caribbean Sea. The Belize<br />

Barrier Reef, an extensive fringing and barrier<br />

reef, developed along an escarpment that abruptly<br />

terminates the 250 km-long Belize continental shelf;<br />

the sea oor plunges to over 183 m (100 fathoms)<br />

just beyond the reef crest (Fig. 1).<br />

In this tropical area of the Caribbean, seasons are<br />

less de ned by temperature and more by rainfall.<br />

The average air temperature ranges from 24 C in<br />

November<strong>–</strong>January, to 27 C in May<strong>–</strong>September<br />

(Purdy et al., 1975). The dry season extends from<br />

February through May; the rainy season extends<br />

from June through November (corresponding to<br />

the peak probability of hurricanes in July<strong>–</strong>October);<br />

December and January are referred to as the transition<br />

season (Auil, 1998). Average annual rainfall<br />

increases in a north to south direction, with 124 cm<br />

near Chetumal Bay, <strong>17</strong>8 cm at Belize City, and<br />

380 cm near the Gulf of Honduras (Purdy et al.,<br />

1975).<br />

The two sampling locations are approximately<br />

70 km apart (Fig. 1). These locations diVer substantially<br />

in both geophysical characteristics and human<br />

activity, as described in more detail below. The<br />

northern location, Basil Jones, is relatively far from<br />

boat traYc centres and manatee resources (abundant<br />

seagrass beds, freshwater, deep channels).<br />

Both locations could provide shelter from surf<br />

surge, with areas suitable for resting and socializing<br />

with other manatees. Compared to Gallows Reef,<br />

seagrass beds appear sparser near Basil Jones.<br />

Northern location<br />

Basil Jones Cut (Fig. 2a), is a few hundred<br />

metres east of Ambergris Caye (18 5 38 N,<br />

87 52 12 W). Inside the Bacalar Chico National<br />

Park and Marine Reserve, this cut is one of several<br />

dozen small breaks in a 50 km, continuous section<br />

of Belize Barrier Reef that hugs the windward shore<br />

(Purdy et al., 1975). Basil Jones Cut (>3 m) is used<br />

by powerboats travelling to a shrimp hatchery,<br />

about a dozen local residents, and a few shermen<br />

or tour operators. The reef lagoon is narrow<br />

(


(a)<br />

(b)<br />

Manatees on the Belize Barrier Reef<br />

Figure 2. Aerial photographs of (a) the northern survey location at Basil Jones Cut (manatees were<br />

observed resting in the deep water channel, i.e. the darker water in the photo), and (b) the southern<br />

survey location east of the Drowned Cayes (Gallows Reef is located along the right margin of<br />

photograph; mainland Belize is approximately 15 km to the west). Photographs by Jimmie C. Smith.<br />

345


346 C. Self-Sullivan et al.<br />

Caye is a solid landmass that blocks manatee travel<br />

from Belize Barrier Reef to Chetumal Bay, a core<br />

centre of manatee activity (Morales-Vela et al.,<br />

2000). Manatees at Basil Jones Cut could travel to<br />

activity centres via two routes (Fig. 1): (1) about<br />

7 km north via Bacalar Chico, a secluded narrow<br />

canal that connects Chetumal Bay to the Caribbean<br />

Sea along the border between Belize and Mexico, or<br />

(2) about 30 km south, around the southern tip of<br />

Ambergris Caye where San Pedro Town is located<br />

(a highly developed tourist destination).<br />

Southern location<br />

Gallows Reef (Fig. 2b), is a section of Belize Barrier<br />

Reef with two breaks: North Gallows Cut<br />

(<strong>17</strong> 30 32 N, 88 3 4 W) and South Gallows Cut<br />

(<strong>17</strong> 27 25 N, 88 2 <strong>17</strong> W). This central section of<br />

the Belize Barrier Reef is discontinuous, with large<br />

breaks in the reef crest, which provide many connections<br />

between deep water and the Belize Barrier<br />

Reef lagoon. Gallows Reef is about 2 km east of the<br />

Drowned Cayes (Fig. 1), an area of mangrove<br />

islands and associated seagrass beds used by<br />

manatees in both winter and summer (Auil, 1998;<br />

Sullivan et al., 1999; LaCommare et al., 2001).<br />

About 15 km due east of Belize City, this string of<br />

islands provides potential navigational ‘stepping<br />

stones’ from the reef to the Belize River, a longterm<br />

manatee activity centre identi ed from<br />

modern aerial survey data (Auil, 1998) and prehistoric<br />

archaeological data (McKillop, 1984).<br />

Throughout this shallow coastal zone, boat traYc is<br />

frequent, including shing boats, tugboats pulling<br />

sugar barges, recreational boats, tour boats, and<br />

water taxis. English Channel, a deep-water shipping<br />

route into the major port at Belize City, is used by<br />

cargo ships, tankers, and cruise ships. Cruise ships<br />

and sugar ships, which are too large for the port,<br />

have berths within the barrier reef lagoon. Small<br />

fast boats transport tourists in all directions from<br />

the cruise ships and Belize City, including welltravelled<br />

routes through the cuts at Gallows Reef<br />

to TurneVe Atoll. Tugboats tow barges of sugar<br />

from points north to temporary sites within the<br />

Drowned Cayes and then to the sugar ship.<br />

Sampling methods<br />

Based on year-round eVort at Basil Jones Cut,<br />

seasonal periods representing winter (December<br />

through March) and summer (June through<br />

August) were chosen for eYcient allocation of<br />

sampling eVort at Gallows Reef. At Basil Jones<br />

Cut, opportunistic observations of manatees by a<br />

local resident (the second author) began prior to<br />

1995 and extended beyond 1997 (Smith, 2000). A<br />

2-year period (April 1995<strong>–</strong>March 1997) of consistent<br />

eVort was selected for the purpose of the present<br />

analysis (Fig. 3). Preliminary studies indicated that<br />

manatees were not present in December-February;<br />

hence, more eVort was allocated to summer<br />

months. To determine whether manatees from Basil<br />

Jones Cut were re-sighted further south, surveys<br />

were extended to Gallows Reef during a study of<br />

the Drowned Cayes by the primary author (Sullivan<br />

et al., 1999). Sampling eVort was limited to winter<br />

(December<strong>–</strong>March) and summer (June<strong>–</strong>August) at<br />

Gallows Reef (1999<strong>–</strong>2001).<br />

Observation and recording procedures were<br />

similar at both locations, following a protocol that<br />

minimized disturbance by in-water observers<br />

(Smith, 2000). Each sample consisted of a 20-min<br />

continuous scan (Lehner, 1996) around a xed<br />

survey point. One to two snorkellers continuously<br />

scanned 360 around the survey point, while oating<br />

at the water’s surface. All samples were collected<br />

between 0800 and 1600 h local time. Only one<br />

sample was taken at each survey point on any given<br />

day at Gallows Reef; no more than two samples<br />

(one in the morning about 1000 h and one in the<br />

afternoon about 1600 h) were taken at the survey<br />

point at Basil Jones Cut.<br />

To determine sighting probability, any manatee<br />

observed during the scan, was recorded as a<br />

‘sighting’. If a manatee approached after the end of<br />

the scan, it was not recorded as a sighting although<br />

it could have been photographed, if feasible.<br />

‘Group size’ was recorded as the total number of<br />

individual manatees present in a scan. In other<br />

words the number of sightings within a 20-min scan<br />

was limited (0, 1); group size was unlimited (0, 1, 2,<br />

3, . . ., N). Group sizes greater than 1 were recorded<br />

only if more than one manatee was observed<br />

simultaneously or if sequential observations were of<br />

uniquely marked individuals. Sketches and photographs<br />

were used to record individual identities at<br />

Basil Jones Cut; photographs and video tapes were<br />

used to record individual identities at Gallows<br />

Reef. When visibility was too poor for positive<br />

identi cation of an individual, it was recorded as an<br />

‘unknown’.<br />

Behaviours (resting, feeding, socializing, milling,<br />

and travelling), body size (calf, adult) and gender<br />

(male, female) were recorded when possible.<br />

De nitions of behavioural activities were: (a) when<br />

‘resting’, the manatee was stationary, either in contact<br />

with the sea oor or at mid-water, occasionally<br />

rising in the vertical direction for breaths, but with<br />

no horizontal movement, no rooting or chewing,<br />

and no reaction to observer, (b) when ‘feeding’, the<br />

manatee was rooting or chewing in a seagrass bed<br />

or had seagrass parts trailing from it’s mouth when<br />

it rose above the bottom, (c) when ‘socializing’, one<br />

manatee touched or followed another, (d) when<br />

‘milling’, the direction of movement changed both<br />

vertically and horizontally with no consistent orientation<br />

to other manatees, to food, or in any one


direction, and (e) when ‘travelling’, the manatee was<br />

moving horizontally in one consistent direction,<br />

either towards or away from the survey point.<br />

To aid in interpretation of results at Gallows<br />

Reef, additional data were collected regarding the<br />

context of samples. As a check on visibility bias,<br />

assistants on a boat (anchored at the survey point)<br />

recorded surface behaviours of manatees relative to<br />

the in-water observer(s). Environmental measurements<br />

collected immediately following the sample<br />

included: (1) sea-surface water temperature using a<br />

thermometer (analogue or digital), (2) sea-surface<br />

salinity using a refractometer, and (3) vertical<br />

visibility using an eight-inch Secchi disk. In no<br />

instance was a manatee sighted by above-water<br />

assistants that was not also sighted by the in-water<br />

observer(s).<br />

The analyses were designed to account for diVerences<br />

in sampling eVort at Basil Jones (n=336) and<br />

Gallows Reef (n=45). Independent variables were<br />

season (winter, summer) at Gallows Reef and<br />

month (January through December) at Basil Jones.<br />

Dependent variables at both locations included: (a)<br />

group size, (b) frequency of surveys with manatees<br />

present or absent. Environmental measures of sea<br />

Manatees on the Belize Barrier Reef<br />

Figure 3. Individual manatee sightings at Basil Jones (April 1995<strong>–</strong>March 1997). Black boxes indicate the manatee was<br />

sighted at least once during the month; seasonal code indicates dry (white), rainy (dark grey), and transitional (light grey)<br />

months. Monthly probability of sighting (S/T) is de ned as the number of surveys that manatees were present, divided<br />

by the total number of surveys for each month. Note: BJ14 (an unmarked male) is included with unknowns in this gure.<br />

surface temperature, salinity, and visibility were<br />

also analysed at Gallows Reef. Non-parametric<br />

statistical tests of continuous variables included<br />

the Mann<strong>–</strong>Whitney U, and the Kruskal<strong>–</strong>Wallis<br />

(Lehner, 1996). Contingencies were tested using the<br />

Fisher’s exact test and the Freeman<strong>–</strong>Tukey deviate<br />

(Bishop et al., 1975).<br />

Results<br />

347<br />

Manatees were documented at both locations on<br />

the Belize Barrier Reef during the summer months,<br />

but not during the winter months. On 40% of all<br />

surveys, at least one manatee was sighted. As<br />

follows, analyses were speci c to each location.<br />

Basil Jones Cut<br />

At least one manatee was sighted on 42% of the 336<br />

surveys at Basil Jones Cut (Fig. 3), and mean group<br />

size varied signi cantly among survey months<br />

(Kruskal<strong>–</strong>Wallis H=119, df=6, P


348 C. Self-Sullivan et al.<br />

Figure 4. At Basil Jones, group size varied signi cantly by<br />

month (1995<strong>–</strong>1997). Horizontal bars indicate means and<br />

circles indicate the range of values.<br />

were present at Basil Jones during the rainy season<br />

(June through November) and absent from<br />

December through March (Fig. 3). They returned<br />

at the end of the dry season (April/May).<br />

Manatees stayed in the calm, deep water of the<br />

channel inside the reef crest. The primary activity<br />

was resting, with occasional socializing. During one<br />

sighting two manatees were observed with seagrass<br />

trailing from their mouths. During another sighting,<br />

two manatees were observed rooting in an area<br />

of sparse seagrass approximately 20 m east of the<br />

resting area. On one occasion, a bull shark was<br />

observed with the manatees, with no interaction.<br />

Travel into the shallower areas of the lagoon system<br />

was rare. Trends in direction of travel by manatees<br />

outside the reef crest could not be determined.<br />

For 87% of the manatees sighted, identity was<br />

determined (Fig. 3). Seventeen individual manatees<br />

had unique markings and were documented by<br />

sketches; most were photographed at least once.<br />

Fifteen of these individuals were observed to be<br />

males. No calves or females were sighted; the sex of<br />

only two identi able individuals was undetermined.<br />

The sex of unknowns (13% of the sightings for<br />

which individual identity could not be determined)<br />

was undetermined in most cases.<br />

Group composition was uid, with no detectable<br />

long-term associations among individuals (Fig. 3).<br />

Two males (BJ01 and BJ08) were re-sighted in each<br />

of 3 years sampled. All seven males that were rst<br />

identi ed in 1995 were re-sighted the next year. Ten<br />

individuals, eight males and two undetermined,<br />

were newly identi ed in 1996. The pattern of<br />

re-sightings within each year was variable; some<br />

individuals were present only one month each<br />

season, others departed and returned after a break<br />

Figure 5. At Gallows Reef, manatees were absent on all<br />

18 winter surveys and present on 12 of 27 summer<br />

surveys.<br />

of 1<strong>–</strong>3 months. One individual ‘White Patch’ (BJ01)<br />

came and went regularly (1994<strong>–</strong>1997), being sighted<br />

<strong>17</strong> times over seven consecutive months in 1996.<br />

The most frequently sighted individual (BJ03) was<br />

present in 41 surveys between April and August<br />

1996; however, he was not re-sighted again until<br />

June of 1998.<br />

Gallows Reef<br />

At least one manatee was sighted on 27% of the 45<br />

surveys at Gallows Reef (Fig. 5). Presence diVered<br />

signi cantly between seasons (Fisher’s exact<br />

phi=0.492, df=1, P=0.001). Manatees were absent<br />

on 100% of surveys during the winter season<br />

(Freeman<strong>–</strong>Tukey deviate z= "3.494) and present<br />

during 44% of the surveys in the summer season<br />

(Freeman<strong>–</strong>Tukey deviate z=1.611). Group size<br />

ranged from one to three, with the larger groups<br />

occurring only in late July and August. ‘White<br />

Patch’ (BJ01) who was frequently re-sighted at Basil<br />

Jones, was re-sighted at Gallows Reef in 1999.<br />

Sex was determined to be male for 10 of the <strong>17</strong><br />

manatees observed during all sightings. No calves<br />

or females were sighted, although the sex of seven<br />

manatees was undetermined.<br />

In general, manatees approached the in-water<br />

observer, paused momentarily and then retreated<br />

beyond visible range, remaining within view of<br />

assistants in the boat. On several occasions, the<br />

same manatee approached the in-water observer<br />

and retreated multiple times during a 20-min scan.<br />

On two occasions, more than one identi able<br />

manatee approached the in-water observer, both<br />

simultaneously and sequentially. Only once did a<br />

second individual approach the in-water observer<br />

after the end of the scan. The primary activity<br />

was milling, with occasional socializing. Rarely<br />

manatees were observed with seagrass trailing from<br />

their mouths; however, they were not observed<br />

rooting into the substrate for rhizomes.


Season signi cantly aVected mean sea surface<br />

temperature (Mann<strong>–</strong>Whitney U=33, n=18, 27,<br />

P20 C) such as natural springs and<br />

human-made attractions such as the warm-water<br />

eZuents of power plants (Packard & Wetterqvist<br />

1986). However, the concept of an essential area is<br />

open to critique based on more recent studies<br />

of individual manatee movements using satellite<br />

telemetry. Individual manatees from the Florida<br />

population vary widely in seasonal movements<br />

(Deutsch et al., 2000). Some travel long distances<br />

along the east coast of the USA, others travel short<br />

distances within Florida or between Florida and<br />

Georgia, and still others appear to be year-round<br />

residents remaining in one activity centre. Since<br />

some southern-most ranges overlap with other<br />

northern-most ranges of Florida manatees (T. m.<br />

latirostris), factors other than ambient water<br />

temperature appear to interact in determining<br />

seasonal movements. Perhaps ‘seasonal activity<br />

centre’ would be a better term for resident<br />

manatees.<br />

We hypothesize that access to freshwater is more<br />

of a directive factor than temperature in in uencing<br />

seasonal manatee use of the reef in the Belize<br />

coastal zone. Alternatively, individual manatee<br />

movements may be determined by a complex interaction<br />

of many factors experienced during a lifetime,<br />

including learning processes in uencing how<br />

travel routes are stored and retrieved from memory.<br />

Our reasoning is as follows.<br />

Even though manatee presence/absence on the<br />

reef was associated with water temperature, the<br />

same variation in seasonal water temperature was<br />

found in the adjacent Drowned Cayes where<br />

manatees were observed year-round during the<br />

same sampling period (Sullivan et al., 1999;<br />

LaCommare et al., 2001). Water temperature in the<br />

study area ranged between 25 C in the winter<br />

and 31 C in the summer, well above the incipient<br />

lethal level (as de ned by Fry, 1947) of cold<br />

tolerance for manatees (20 C c.f. Irvine, 1983).<br />

Temperature may have been correlated with<br />

another, undetermined, directive factor or gradient.<br />

An alternative hypothesis might be that manatees<br />

move further from estuaries during the summer<br />

rainy season when freshwater plumes from rivers<br />

are more likely to penetrate further into the coastal<br />

zone, meaning that manatees are more likely to be


350 C. Self-Sullivan et al.<br />

at the reef in the summer. Annual salinity range at<br />

Gallows Reef was 34.5<strong>–</strong>38.0 ppt. Osmoregulation<br />

experiments on Florida manatees indicate that: (1)<br />

they are good osmoregulators in both fresh (0‰)<br />

and marine (34‰) environments (Ortiz et al., 1998);<br />

and (2) although they drink large amounts of<br />

freshwater when available, they do not drink<br />

marine water, but possibly oxidize fat to meet their<br />

water needs when restricted to eating seagrass in the<br />

marine environment (Ortiz et al., 1999). Manatees<br />

are considered to be dependent on freshwater in<br />

Florida (Hartman, 1979), and periodic access to<br />

freshwater is thought to be important to manatees<br />

in Belize (Gibson, 1995). However, it is not known<br />

whether nor how often manatees might move from<br />

the oVshore activity centres to mainland sources of<br />

freshwater in Belize. Although this distance is all<br />

well within 1-day’s travel range, manatees within<br />

the Drowned Cayes area are often sighted with<br />

dozens of salt-water barnacles covering their bodies<br />

(Self-Sullivan, unpublished data), an indication<br />

of long periods of time spent in the marine<br />

environment (Husar, 1997).<br />

Alternatively, underwater springs near our study<br />

locations may dry up or become unpalatable during<br />

winter. One manatee, sighted three times at Basil<br />

Jones in January and February 2001, was near a<br />

spring adjacent to the channel (G. Smith, unpublished<br />

data). Whether the manatee was drinking<br />

from the underwater spring could not be determined;<br />

the water was sulphurous as determined by<br />

smell and colour. Since we documented higher<br />

surface salinity in the summer compared to the<br />

winter at Gallows Reef, we do not believe that<br />

rainfall provides a lens of fresh surface water at the<br />

reef. Possibly the higher salinity was related to<br />

evaporation during summer months.<br />

Another hypothesis might be that the breaks in<br />

the reef could serve as a stopover site for manatees<br />

travelling in search of fresh water, in an east<strong>–</strong>west<br />

direction to and from oVshore atolls. Gallows Reef<br />

is midway between the Belize River (a fresh<br />

water source) and TurneVe Atoll, located further<br />

oVshore. Manatees have been opportunistically<br />

sighted at TurneVe Atoll, a large complex mangrove<br />

island-seagrass-coral reef system similar to<br />

the Drowned Cayes (Auil, 1998; Barbara Bilgre,<br />

pers. comm.).<br />

Seasonal distribution also might be related to<br />

winter storms from the northwest. Cold air masses<br />

from North America frequently aVect both temperature<br />

and wind strength during October through<br />

January (Purdy et al., 1975). ‘Northers’, as the local<br />

people call these events, may lower the sea surface<br />

by as much as 0.8 m. These events have a greater<br />

eVect than spring tides (0.5 m) on water depth in<br />

shallow northern lagoons. If manatees move away<br />

from shallow areas inside the reef, then absence<br />

would be more likely after storm events in the<br />

winter.<br />

Alternatively, manatees may have learned that<br />

the reef provides shelter from high surf during the<br />

summer hurricane season. For example, three<br />

manatees were at Gallows Reef 2 days after<br />

Tropical Storm Chantal hit the coast of Belize in<br />

2001 (Self-Sullivan, unpublished data). This is the<br />

largest group sighted at Gallows Reef. If manatees<br />

that move into the protected lagoon system sense<br />

protection from strong currents created by storm<br />

surge, breaks in the reef might serve as stopover<br />

sites for individuals moving back out to the atolls or<br />

continuing their north-south travel outside the<br />

reef. For obvious logistical reasons, such movements<br />

during hurricanes would be better studied by<br />

satellite telemetry than boat-based surveys.<br />

Clearly, there are several physical factors that are<br />

correlated with seasonal changes in the habitat used<br />

by manatees in Belize. To tease out the relative<br />

importance of these factors, we would recommend<br />

a combination of regional studies of individual<br />

manatee movements using satellite telemetry as well<br />

as simultaneous site-speci c studies at breaks in the<br />

reef. This type of approach would also provide a<br />

better understanding of how external environmental<br />

conditions interact with physiological states,<br />

as outlined below.<br />

Relation between travel and seasonal male<br />

reproductive activity<br />

Noting the distinctive absence of females at the<br />

breaks in the reef that we studied, a hypothesis<br />

emerged related to potential seasonal changes in<br />

reproductive activity of males. The evidence that<br />

manatees show seasonal patterns in reproductive<br />

behaviour is still ambiguous. Early reports by<br />

Hartman (1979) and Marsh et al. (1978) found no<br />

evidence for strong seasonality in <strong>Sirenian</strong> reproductive<br />

behaviour. However, Best (1982) reported<br />

seasonal breeding in the Amazonian manatee (T.<br />

inunguis). More recently, seasonal spermatogenesis<br />

has been reported in both Florida manatees<br />

(Hernandez et al., 1995) and dugongs (Dugong<br />

dugon) (Marsh, 1995). Similarly, seasonal variation<br />

in female reproductive hormones has been shown in<br />

captive Florida manatees (Larkin, 2000). However,<br />

seasonal reproductive activities were not tightly<br />

coordinated within populations (Larkin, 2000; Best,<br />

1982; Marsh et al., 1984; Marsh, pers. comm.).<br />

By tracking manatees using VHF radio and<br />

satellite tags since 1997, Powell et al. (2001) found<br />

more variation in male than female movement<br />

patterns in coastal lagoons south of Belize City. For<br />

example, females and some males remained inside<br />

the enclosed estuary year-round. However, a few<br />

males wandered outside the lagoon system, north<br />

and south along the mainland coast of Belize; one


male travelled at least as far as Belize City, 33 km<br />

north (Powell, pers. comm.).<br />

Based on published literature and our observations<br />

of only males on the reef with seasonal<br />

peaks and absences, we hypothesized that male<br />

manatees use the reef as a landmark during searches<br />

for oestrous females throughout the coastal lagoon<br />

system. If there is a winter low in spermatogenesis,<br />

then movement rates might decline. If there is a<br />

peak in both oestrous females and searching behaviour<br />

of males in summer, then male movement rates<br />

may increase. Variations in seasonal reproductive<br />

activity within a population might explain the<br />

diVerence in sighting peaks between Basil Jones<br />

(May<strong>–</strong>June) and Gallows Reef (July<strong>–</strong>August).<br />

Alternatively, sighting peaks could indicate<br />

clustered movements by males along the reef.<br />

To test these alternate hypotheses, we would<br />

recommend satellite-telemetry studies of both males<br />

and females in activity centres adjacent to the reef<br />

(Drowned Cayes and Chetumal Bay), combined<br />

with simultaneous site-speci c studies of individually<br />

identi able manatees. Such studies could<br />

answer the critical question of the degree to which<br />

there is a seasonal peak in reproductive activity<br />

of manatees in Belize and the degree to which<br />

male manatees move along the reef during peak<br />

reproductive periods.<br />

Implications for expanding populations in<br />

fragmented habitat<br />

Based on our sightings of ‘White Patch’ (BJ01) at<br />

both the northern (1994<strong>–</strong>1997) and southern (1999)<br />

sampling locations, we hypothesized manatees<br />

move along the reef in a north-south direction.<br />

However, we were not able to monitor both locations<br />

simultaneously. Extensive studies based on<br />

re-sightings of individually marked manatees have<br />

been successful in Florida (Reid et al., 1991) and<br />

should be continued in Belize. Over the long-term,<br />

these studies can be combined with satellite-tagging<br />

studies to determine variation in movement patterns<br />

during the lifetimes of individuals within a<br />

population (Beck & Reid, 1995; Reid et al., 1995).<br />

By comparing records at TurneVe Atoll (Bilgre,<br />

unpublished data), the Drowned Cayes, Gallows<br />

Reef (Smith, 2000; Self-Sullivan, unpublished data),<br />

and Basil Jones (Smith, 2000), a more accurate<br />

model of what in uences manatee movements in the<br />

centre of the species’ range may emerge for comparison<br />

with what is known from the northern<br />

extreme of the range, as studied in Florida.<br />

Long distance movements in uence gene ow<br />

among subpopulations, as re ected in complex patterns<br />

of genetic variation among manatees in the<br />

greater Caribbean area (Garcia-Rodriguez et al.,<br />

1998). Movements of individual Florida manatees<br />

along the Atlantic coast of North America range<br />

Manatees on the Belize Barrier Reef<br />

351<br />

from 44 km to 2360 km (median=309 km) at rates<br />

of 25 km to 87 km per day (Deutsch et al., 2000). In<br />

comparison, the Belize Barrier Reef extends only<br />

about 220 km, with many breaks that could be used<br />

as stopover sites or for ingress to manatee activity<br />

centres such as Chetumal Bay, Cayes oV Belize<br />

City, Southern Lagoon, the Gulf of Honduras, and<br />

numerous coastal rivers and bays (Auil, 1998).<br />

Considering the long travel range of Florida<br />

manatees, Antillean manatees could move among<br />

population centres along the Yucatan Peninsula<br />

from Mexico through Belize to Guatemala and<br />

Honduras (Auil, 1998). During population expansion<br />

following historical exploitation, manatees<br />

from Belize may colonize unoccupied habitat<br />

between population centres in a wider region, with<br />

positive in uences on gene ow.<br />

The reef may also in uence connectivity within<br />

the Belize Barrier Reef lagoon system, decreasing<br />

the probability of inbreeding by increasing movements<br />

between southern areas and Chetumal Bay.<br />

The deep water just east of the barrier reef oVers an<br />

unobstructed north-south travel corridor and the<br />

reef oVers a physical feature that could be used for<br />

navigation. Use of this travel route by manatees<br />

could be an alternative to movement through the<br />

complex labyrinth created by the mangroveseagrass-coral<br />

lagoon system inside the barrier reef.<br />

An analogy might be a loop highway around a<br />

metropolitan maze of small roads.<br />

Consistent with knowledge of ‘bachelor males’ in<br />

other marine mammals species, we hypothesize that<br />

the turnover of individuals using stopover sites on<br />

the reef could be related to turnover in male access<br />

to breeding females. In 2001, local tour guides and<br />

staV at Bacalar Chico Reserve reported that they<br />

‘regularly’ observed manatees in a side lagoon<br />

adjacent to Bacalar Chico and at another cut in the<br />

reef between Bacalar Chico and Basil Jones (Smith,<br />

unpublished data). Another local tour guide<br />

reported that he observed a group of ve manatees<br />

inside the reef crest at GoV’s Caye (a sand caye on<br />

the Belize Barrier Reef just south of Gallows Reef)<br />

in November 2001 (Self-Sullivan, unpublished<br />

data). And a third tour guide reported seeing a<br />

manatee outside the reef crest at South Water Caye<br />

(a sand caye on the Belize Barrier Reef still further<br />

south) in June 2002 (Self-Sullivan, unpublished<br />

data). We recommend studies to determine if these<br />

are young dispersing males, prime males between<br />

periods of breeding activity, or old males unlikely to<br />

breed again. Answering these questions will provide<br />

insight to how genetic heterogeneity may be maintained<br />

in relatively isolated subpopulations of<br />

manatees.<br />

We recommend that the broader implications of<br />

this highly speci c study should be considered in the<br />

context of eVorts to design marine reserve systems


352 C. Self-Sullivan et al.<br />

within the Caribbean. The concept of ‘connectivity’<br />

was used by Roberts (1997:1454) to draw attention<br />

to how ‘local populations may depend on processes<br />

occurring elsewhere’. Although Roberts referred to<br />

patterns of larval transport among coral reefs, the<br />

concept also should be applied to large-bodied<br />

vulnerable species that use the lagoon systems protected<br />

by reefs e.g., manatees. We use the concept of<br />

‘stopover sites’ in a manner similar to its use in<br />

studies of migratory bird species. For example,<br />

Higuchi et al. (1996) identi ed stopover sites where<br />

migratory cranes paused for less than 10 days<br />

during travel between breeding and non-breeding<br />

locations. Although Antillean manatees clearly are<br />

not migratory as a species, the concept of a stopover<br />

site may be applied to a relatively short<br />

interruption of travel by individuals. Protecting<br />

stopover sites within local lagoon systems could be<br />

as important for conserving genetic heterogeneity,<br />

as protecting long-distance travel routes connecting<br />

fragmented populations on a regional scale.<br />

Conclusions<br />

Developing eVective strategies for conservation of<br />

manatee habitat requires knowledge of activity<br />

centres and movements among them. We documented<br />

manatee presence at breaks in the Belize<br />

Barrier Reef, with signi cantly more sightings in<br />

the summer (rainy) season than in the winter<br />

(transition/dry) season. Groups were small (average<br />

of two manatees), some individuals were repeatedly<br />

sighted and others were not. At least one individual<br />

moved between the northern and southern breaks<br />

that we monitored. All of the manatees for which<br />

sex was determined were male and no calves<br />

were sighted. However, we are cautious about<br />

interpretation of these data, due to the inductive<br />

nature of this study. We proposed several hypotheses<br />

related to physical (temperature, salinity,<br />

depth, storm surge) and physiological (thermoregulation,<br />

osmoregulation, reproductive status)<br />

factors potentially aVecting manatee use of breaks<br />

in the reef. To test these hypotheses, we recommend<br />

simultaneous studies of identi able manatees at<br />

several locations on the reef. In addition, satellitetagging<br />

techniques would be appropriate for determining<br />

the seasonality and extent of individual<br />

manatee movements among activity centres and<br />

stopover sites along travel routes associated with<br />

geophysical characteristics of the Belize Barrier<br />

Reef. Based on results of this study, we propose<br />

that reefs should be considered part of the network<br />

of travel routes and stopover sites identi ed as<br />

important components maintaining connectivity<br />

within manatee habitat in Belize and between fragmented<br />

populations of Antillean manatees in the<br />

Caribbean.<br />

Acknowledgments<br />

This research project was reviewed by the Belize<br />

National Manatee Working Group, Coastal Zone<br />

Management Authority and Institute, and conducted<br />

under Scienti c Research Permits issued by<br />

the Belize Forestry Department, Conservation<br />

Division. Funding and in-kind support were provided<br />

by The Oceanic Society of San Francisco, The<br />

Lerner<strong>–</strong>Gray Marine Science Research Fund, The<br />

Earthwatch Institute, Spanish Bay Resort, Texas<br />

A&M University, University of Massachusetts-<br />

Boston, and the National Science Foundation<br />

Graduate Fellowship Program. We are extremely<br />

grateful to our Belizean eld assistants Pach<br />

Muñoz, Landy Requena, Jerry Requena, and<br />

Gilroy Robinson; our Belizean student interns<br />

Maxine Monsanto, Seleem Chan, and Clifton<br />

Williams; our logistics interns Liz Johnstone<br />

and Pam Quayle; and our Oceanic Society and<br />

Earthwatch Institute volunteers for their assistance<br />

in data collection. We thank Jimmie C. Smith for<br />

providing aerial photographs of our study area.<br />

Our gratitude is extended to Nicole Auil, Miriam<br />

Marmontel, and Leon David Olivera-Gomez for<br />

comments on previous drafts of this manuscript.<br />

Literature Cited<br />

Auil, N. (1998) Belize Manatee Recovery Plan. UNDP/<br />

GEF Coastal Zone Management Project, BZE/92/G31,<br />

Belize/UNEP Caribbean Environment Programme,<br />

Kingston.<br />

Beck, C. A., & Reid, J. P. (1995) An automated photoidenti<br />

cation catalog for studies of the life history of<br />

the Florida manatee. In: T. J. O’Shea, B. B. Ackerman<br />

& H. F. Percival (eds.) Population Biology of the Florida<br />

Manatee, pp. 120<strong>–</strong>123 National Biological Service,<br />

Washington, D.C.<br />

Bengtson, J. L. & Magor, D. (1979) A survey of manatees<br />

in Belize. Journal of Mammalogy 60, 230<strong>–</strong>232.<br />

Best, R. C. (1982) Seasonal Breeding in the Amazonian<br />

Manatee, Trichechus inunguis (Mammalia: Sirenia).<br />

Biotropica 14, 76<strong>–</strong>78.<br />

Bishop, Y. M. M., Fienberg, S. E. & Holland, P. W.<br />

(1975) Discrete Multivariate Analysis. The MIT Press,<br />

Cambridge.<br />

CEP/UNEP. (1995) Regional Management Plan for<br />

the West Indian Manatee, Trichechus manatus.<br />

CEP Technical Report No. 35. UNEP Caribbean<br />

Environment Programme, Kingston.<br />

Deutsch, C. J., Reid, J. P. Bonde, R. K., Easton, D. E.,<br />

Kochman, H. I. & O’Shea, T. J. (2000) Seasonal<br />

movements, migratory behavior, and site delity of<br />

West Indian manatees along the Atlantic coast of the<br />

United States as determined by radio-telemetry. Final<br />

Report. Research Work Order No. 163. Florida<br />

Cooperative Fish and Wildlife Research Unit,<br />

U.S. Geological Survey and University of Florida,<br />

Gainesville.<br />

Fry, F. E. J. (1947) EVects of the environment on animal<br />

activity. University of Toronto Studies Biological Series


55. Ontario Fisheries Research Laboratory 68.<br />

University of Toronto Press, Toronto.<br />

Garcia-Rodrieguez, A. I., Bowen, B. W., Domning, D. P.,<br />

Mignucci-Giannoni, A. A., Marmontel, M., Montoya-<br />

Ospina, R. A., Morales-Vela, B., Rudin, M., Bonde,<br />

R. K. & McGuire, P. M. (1998) Phylogeography of the<br />

West Indian manatee (Trichechus manatus): how many<br />

populations and how many taxa? Molecular Ecology 7,<br />

1137<strong>–</strong>1149.<br />

Gibson, J. (1995) Managing Manatees in Belize. M.S.<br />

Thesis. Department of Marine Sciences and Coastal<br />

Management, University of Newcastle upon Tyne.<br />

Hartman, D. S. (1979) Ecology and Behavior of<br />

the Manatee (Trichechus manatus) in Florida.<br />

Special Publication No. 5, The American Society of<br />

Mammalogists.<br />

Hernandez, P., Reynolds, J. E., Marsh, H. & Marmontel,<br />

M. (1995) Age and seasonality in spermatogenesis of<br />

Florida manatees. In: T. J. O’Shea, B. B. Ackerman &<br />

H. F. Percival (eds.) Population Biology of the Florida<br />

Manatee, pp. 84<strong>–</strong>97 National Biological Service,<br />

Washington, D.C.<br />

Higuchi H., Ozaki K., Fijita G., Minton J., Ueta M.,<br />

Soma M., & Mita N. (1996) Satellite tracking of<br />

White-naped crane migration and the importance of the<br />

Korean demilitarized zone. Conservation Biology 10,<br />

806<strong>–</strong>812.<br />

Hilton-Taylor, C. (compiler). (2001) 2000 IUCN Red List<br />

of Threatened Species. IUCN, Gland, Switzerland and<br />

Cambridge, UK.<br />

Husar, S. L. (1977) The West Indian manatee (Trichechus<br />

manatus). Wildlife Research Report 7. U.S.<br />

Department of the Interior, Fish and Wildlife Service,<br />

Washington, D. C.<br />

Irvine, A. B., (1983) Manatee metabolism and its in uence<br />

on distribution in Florida. Biological Conservation 25,<br />

315<strong>–</strong>334.<br />

Koelsch, J. K. (1997) The seasonal occurrence and ecology<br />

of Florida manatees (Trichechus manatus<br />

latirostris) in coastal waters near Sarasota, Florida.<br />

Master’s Thesis, Department of Biology, University of<br />

South Florida.<br />

LaCommare, K. S., Sullivan, C. S. & Brault, S. (2001)<br />

Distribution and foraging ecology of Antillean<br />

manatees (Trichechus manatus) in the Drowned Cays<br />

area of Belize, Central America. Abstracts, 14th<br />

Biennial Conference on the Biology of Marine<br />

Mammals, November 29<strong>–</strong>December 3, 2001,<br />

Vancouver, Canada.<br />

Larkin, I. L. V. (2000) Reproductive endocrinology of the<br />

Florida manatee (Trichechus manatus latirostris):<br />

estrous cycles, seasonal patterns and behavior. Ph.D.<br />

Dissertation. University of Florida.<br />

Lefebvre, L. W., Marmontel, M., Reid, J. P., Rathbun,<br />

G. B. & Domning, D. P. (2001) Status and biogeography<br />

of the West Indian manatee. In: Charles A. Woods<br />

& Florence E. Sergile (eds.) Biogeography of the West<br />

Indies: Patterns and Perspectives, pp. 425<strong>–</strong>474. CRC<br />

Press: New York.<br />

Lehner, P. N. (1996) Handbook of Ethological Methods,<br />

2nd edition. Cambridge University Press.<br />

Marsh, H. (1995) The life history, pattern of breeding, and<br />

population dynamics of the dugong. In: T. J. O’Shea,<br />

Manatees on the Belize Barrier Reef<br />

353<br />

B. B. Ackerman & H. F. Percival (eds.) Population<br />

Biology of the Florida Manatee, pp. 75<strong>–</strong>83. National<br />

Biological Service, Washington, D.C.<br />

Marsh, H., Heinsohn, G. E. & Glover, T. D. (1984)<br />

Changes in the male reproductive organs of the<br />

dugong, Dugong dugon (Sirenia: Dugondidae) with age<br />

and reproductive activity. Australian Journal of<br />

Zoology 32, 721<strong>–</strong>742.<br />

Marsh, H., Spain, A. V. & Heinsohn, G. E. (1978)<br />

Minireview: physiology of the dugong. Comparative<br />

Biochemisty and Physiology 61A, 159<strong>–</strong>168.<br />

McKillop, H. I. (1984) Prehistoric Maya Reliance on<br />

Marine Resources: Analysis of a Midden from Moho<br />

Cay. Belize Journal of Field Archaeology 11, 25<strong>–</strong>35.<br />

Morales-Vela, B., Olivera-Gomez, D., Reynolds, J. E. &<br />

Rathbun, G. B. (2000) Distribution and habitat use by<br />

manatees (Trichechus manatus manatus) in Belize and<br />

Chetumal Bay, Mexico. Biological Conservation 95,<br />

67<strong>–</strong>75.<br />

Ortiz, R. M. Worthy, G. A. J. & Byers, F. M. (1999)<br />

Estimation of water turnover rates of captive West<br />

Indian manatees (Trichechus manatus) held in fresh and<br />

salt water. Journal of Experimental Biology 202, 33<strong>–</strong>38.<br />

Ortiz, R. M., Worthy, G. A. J. & MacKenzie, D. S. (1998)<br />

Osmoregulation in wild and captive West Indian<br />

manatees (Trichechus manatus). Physiological Zoology<br />

71, 449<strong>–</strong>457.<br />

O’Shea, T. J. & Salisbury, C. A. (1991) Belize - a last<br />

stronghold for manatees in the Caribbean. Oryx 25,<br />

156<strong>–</strong>164.<br />

Packard, J. M. & Wetterqvist, O. F. (1986) Evaluation of<br />

manatee habitat systems on the northwestern Florida<br />

coast. Coastal Zone Management Journal 14, 279<strong>–</strong>310.<br />

Powell, J. A., Bonde, R., Aguirre, A. A., Koontz, C.,<br />

Gough, M. & Auil, N. (2001) Biology and movements<br />

of manatees in Southern Lagoon, Belize. Abstract in<br />

Proceedings of the 14th Biennial Conference on the<br />

Biology of Marine Mammals, Vancouver.<br />

Purdy, E. G., Pusey, W. C., III & Wantland, K. F. (1975)<br />

Continental shelf of Belize <strong>–</strong> regional shelf attributes.<br />

In: K. F. Wantland & W. C. Pusey, III (eds.) Studies in<br />

Geology No. 2: Belize Shelf <strong>–</strong> Carbonate Sediments,<br />

Clastic Sediments, and Ecology and a paper on Petrology<br />

and Diagenesis of Carbonate Eolianites of Northeastern<br />

Yucatan Peninsula, Mexico, pp. 1<strong>–</strong>52. The American<br />

Association of Petroleum Geologists, Tulsa.<br />

Reid, J. P., Rathbun, G. B. & Wilcox, J. R. (1991)<br />

Distribution patterns of individually identi able<br />

West Indian manatees (Trichechus manatus) in Florida.<br />

Marine Mammal Science 7, 180<strong>–</strong>190.<br />

Reid, J. P., Bonde, R. K. & O’Shea, T. J. (1995)<br />

Reproduction and mortality of radio-tagged and recognizable<br />

manatees on the Atlantic Coast of Florida. In:<br />

T. J. O’Shea, B. B. Ackerman & H. F. Percival (eds.)<br />

Population Biology of the Florida manatee, pp. <strong>17</strong>1<strong>–</strong>191.<br />

National Biological Service, Washington, D.C.<br />

Roberts, C. M. (1997) Connectivity and management of<br />

Caribbean coral reefs. Science 278, 1454<strong>–</strong>1457.<br />

Sanderson, G. C. (1966) The study of mammal<br />

movements <strong>–</strong> a review. Journal of Wildlife Management<br />

30, 215<strong>–</strong>235.<br />

Sullivan, C. S., Packard, J. M. & Evans, W. E. (1999)<br />

Spring distribution and behavior of Antillean manatees<br />

(Trichechus manatus manatus) in the Drowned Cayes,


354 C. Self-Sullivan et al.<br />

Belize. Abstracts, 13th Biennial Conference on the<br />

Biology of Marine Mammals, November 28<strong>–</strong>December<br />

3, 1999, Wailea, Maui, Hawaii.<br />

Smith, G. W. (2000) Identi cation of Individual Manatee<br />

In the Basil Jones Area of the Bacalar Chico Marine<br />

Reserve and the Drowned Cays Area of Belize. Annual<br />

Report to the National Manatee Working Group,<br />

Coastal Zone Management Institute & Authority,<br />

Belize City.<br />

Weeks, P. & Packard, J. M. (1997) Acceptance of<br />

scienti c management by natural resource dependent<br />

communities. Conservation Biology 11, 236<strong>–</strong>245.


NEWS OF THE WEEK<br />

MEETINGBRIEFS>><br />

1070<br />

AAAS ANNUAL MEETING | 18 TO 22 FEBRUARY 2010 | SAN DIEGO, CALIFORNIA<br />

A symposium organized at the last minute at<br />

the annual meeting of the American Association<br />

for the Advancement of Science (the<br />

publisher of Science) by two of the world’s<br />

most prominent scientific organizations<br />

addressed recent attacks on an increasingly<br />

beleaguered climate science community. The<br />

panel met in the uncertain aftermath of the<br />

stolen e-mails affair and critiques of the<br />

Intergovernmental Panel on Climate Change<br />

(IPCC) (Science, 12 February, p. 768).<br />

The symposium was convened by U.S.<br />

National Academy of Sciences President<br />

Ralph Cicerone, in conjunction with AAAS, at<br />

a time when flaws in the latest<br />

IPCC report, and even the legitimacy<br />

of climate science, have<br />

made headlines. E-mails<br />

uncovered late last year<br />

revealed instances of scientists<br />

on the panel discussing<br />

withholding data and documents<br />

from those with opposing<br />

views, conspiring to keep<br />

contradictory papers out of<br />

influential reports, and<br />

encouraging colleagues to<br />

delete e-mails.<br />

Despite a drumbeat of studies that corroborate<br />

the conclusion that the planet is warming<br />

and human activities are largely responsible,<br />

these recent skirmishes “have really shaken<br />

the confidence of the public in the conduct of<br />

science [overall],” said Cicerone, citing a number<br />

of recent polls on the public perception of<br />

science. “The situation is completely out of<br />

hand,” said climate scientist Gerald North of<br />

Not so fast. Critics have hit IPCC over a false assertion<br />

that Himalayan glaciers would melt away by 2035.<br />

Scientists Grapple With ‘Completely<br />

Out of Hand’ Attacks on Climate Science<br />

E-mail etiquette. Gerald North<br />

says scientists should not sound<br />

like bloggers.<br />

Texas A&M University in College Station,<br />

who has served as an IPCC reviewer. “One<br />

guy e-mailed me to say I’m a ‘whore for the<br />

global warming crowd.’ ” His PowerPoint<br />

presentation at the meeting included a slide<br />

quoting conservative talk show host Glenn<br />

Beck, who suggested that scientists commit<br />

“hara-kari” to atone. “Scientists cannot use<br />

the same tone and rhetorical style as commentators<br />

and bloggers,” North said.<br />

Although much of the session at the<br />

meeting, titled “Ensuring the Transparency<br />

and Integrity of Scientific Research,” focused<br />

on what Harvard University oceanographer<br />

and former AAAS head James<br />

McCarthy called the “abominable”<br />

press coverage, scientists<br />

owned up to their share of<br />

the blame. Small errors in the<br />

2007 report were “careless,”<br />

said McCarthy, but IPCC<br />

should have done a full and<br />

public examination to describe<br />

how they had come about:<br />

“The names of the authors,<br />

who was on the review, what<br />

happened—it all should have<br />

been up there, and it wasn’t<br />

done. And I think that the institution was hurt<br />

as a result,” he said.<br />

The community allowed “the situation to<br />

get out of control,” said Sheila Jasanoff of<br />

Harvard University. She said in general scientists<br />

had to connect better to the public.<br />

“There is a kind of arrogance—we are scientists<br />

and we know best,” Jasanoff said.<br />

“That needs to change.” <strong>–</strong>ELI KINTISCH<br />

26 FEBRUARY 2010 VOL 327 SCIENCE www.sciencemag.org<br />

Published by AAAS<br />

The Latest on<br />

Geoengineering<br />

Preliminary findings presented here suggest<br />

that some proposed techniques to cool the<br />

planet manually may have fewer barriers<br />

than previously thought. But many technical<br />

and societal barriers remain.<br />

Even before they got to the sessions, the<br />

scientists had to contend with a smattering of<br />

activists with drums, cameras, and a megaphone<br />

alleging that the government is already<br />

performing geoengineering through the<br />

spraying of particles, in so-called chemtrails.<br />

Physicist David Keith of the University<br />

of Calgary in Canada addressed the concept<br />

of spreading aerosol droplets in the stratosphere,<br />

where they could block a small fraction<br />

of the sun’s rays. A paper published last<br />

year in Environmental Research Letters suggested<br />

that the leading proposal, spraying<br />

Is a Dolphin a Person?<br />

Are dolphins as smart as people? And if<br />

so, shouldn’t we be treating them a bit better?<br />

Those were the questions scientists<br />

and philosophers debated at a session here<br />

on Sunday.<br />

Dolphins, it turns out, are pretty darn smart.<br />

Panelist Lori Marino, an expert on cetacean<br />

neuroanatomy at Emory University in Atlanta,<br />

said they may be Earth’s second smartest creature,<br />

after humans, of course.<br />

Bottlenose dolphins have bigger brains than<br />

humans (1600 grams versus 1300 grams), and<br />

they have a brain-to-body-weight ratio greater<br />

than that of great apes (but smaller than that of<br />

humans), said Marino. “They are the second<br />

most encephalized beings on the planet.”<br />

But it’s not just size that matters. Dolphins<br />

also have a very complex neocortex, the part<br />

of the brain responsible for problem solving,<br />

self-awareness, and various other traits we<br />

associate with human intelligence. And<br />

researchers have found spindle neurons in<br />

dolphin brains called von Economo neurons<br />

that in humans and apes have been linked to<br />

emotions, social cognition, and even theory of<br />

mind: the ability to sense what others are<br />

thinking. Overall, said Marino, “dolphin<br />

brains stack up quite well to human brains.”<br />

What dolphins do with their brains is also<br />

impressive. Cognitive psychologist Diana<br />

CREDITS (TOP TO BOTTOM): PHOTOS.COM; CONSERVATION HISTORY ASSOCIATION OF TEXAS<br />

on March 11, 2010<br />

www.sciencemag.org<br />

Downloaded from


CREDIT: PHOTOS.COM<br />

sulfur dioxide gas, wouldn’t work. Sulfur<br />

dioxide is converted in the atmosphere into<br />

droplets of sulfuric acid, which would clump<br />

and fall out of the sky before they could have<br />

much cooling effect. To get around this<br />

problem, Keith and colleagues have proposed<br />

using airplanes to spray droplets of<br />

the acid itself, rather than sulfur dioxide. In<br />

unpublished data, the team found that injecting<br />

only “a few megatons per year” of sulfuric<br />

acid could be more than twice as effective<br />

at blocking radiation as starting with<br />

sulfur dioxide.<br />

While scientists are finding ways to overcome<br />

the engineering challenges, the environmental<br />

effects of planet-hacking techniques<br />

remain uncertain. One challenge in geoengineering<br />

a warmed planet is simultaneously<br />

restoring temperatures while minimizing disruption<br />

of rain and precipitation. (Stratospheric<br />

particles lower the total amount of energy striking<br />

Earth, the driver of precipitation.)<br />

In previous modeling efforts, adding<br />

sun-blocking particles uniformly across the<br />

globe has tended to undercool the poles<br />

while overcooling the equator. So Kenneth<br />

Reiss of Hunter College of the City University<br />

of New York has been working with dolphins in<br />

aquariums for most of her career, and she said<br />

their social intelligence rivals that of the great<br />

apes. Dolphins can recognize themselves in a<br />

mirror, a sign of self-awareness. They can<br />

understand complex gesture “sentences” from<br />

humans. And they can learn to poke an underwater<br />

keyboard to request toys. “Much of their<br />

learning is similar to what we see with young<br />

children,” said Reiss.<br />

So if dolphins are so similar to people,<br />

shouldn’t we be treating them more like people?<br />

“The very traits that make dolphins interesting<br />

to study,” said Marino, “make confining<br />

them in captivity unethical.” She noted, for<br />

example, that, in the wild, dolphins have a<br />

home range of about 100 square kilometers. In<br />

captivity, they roam one 10-thousandth of one<br />

percent of this area.<br />

Far worse, Reiss said, is the massive dolphin<br />

culling ongoing in some parts of the world,<br />

which she documented with a graphic video of<br />

dolphins being drowned and stabbed in places<br />

like the Japanese town of Taiji.<br />

Thomas White, a philosopher at Loyola<br />

Marymount University in Redondo Beach,<br />

California, suggested that dolphins aren’t<br />

merely like people—they may actually be people,<br />

or at least, “nonhuman persons.” Defining<br />

exactly what it means to be a person is difficult,<br />

White said, but dolphins seem to fit the check-<br />

Caldeira, a climate scientist at the Carnegie<br />

Institution for Science in Stanford, California,<br />

modeled various approaches to try to<br />

counteract a severe warming—the result of a<br />

doubling of preindustrial CO 2 concentration.<br />

In work yet to be published, he distributed<br />

the particles unevenly to try to minimize<br />

those effects; for example, by putting<br />

more at the poles versus the equator. (Global<br />

warming is greatest in the Arctic.) In models,<br />

that strategy helped fix the undercooling/overcooling<br />

problem, but it worsened<br />

the effects on precipitation. “There’s a<br />

complex problem of how do you balance<br />

the damage that you do against the benefit,”<br />

said Caldeira.<br />

That said, simulating either geoengineering<br />

approach to counteract global warming—<br />

distributing particles globally or focusing on<br />

the poles—suggests a cooler world with less<br />

disruption of rain patterns than one in which<br />

warming continues unabated. “In a highglobal-warming<br />

world, more people would be<br />

better off with geoengineering, but some people<br />

would be worse off,” he said.<br />

<strong>–</strong>ELI KINTISCH<br />

Social smarts. Dolphins display many of the<br />

same behaviors humans do.<br />

list many philosophers agree on. There are the<br />

obvious ones: They’re alive, aware of their environment,<br />

and have emotions; but they also<br />

seem to have personalities, exhibit self-control,<br />

and treat others appropriately, even ethically.<br />

When it comes to what defines a person, said<br />

White, “dolphins fit the bill.”<br />

Still, experts caution that the scientific<br />

case for dolphin intelligence is based on relatively<br />

little data. “It’s a pretty story, but it’s<br />

very speculative,” says Jacopo Annese, a<br />

neuroanatomist at the University of California,<br />

San Diego. Despite a long history of research,<br />

scientists still don’t agree on the roots of intelligence<br />

in the human brain, he says. “We don’t<br />

know, even in humans, the relationship<br />

between brain structure and function, let alone<br />

intelligence.” And, Annese says, far less is<br />

known about dolphins. <strong>–</strong>DAVID GRIMM<br />

With reporting by Greg Miller.<br />

More Highlights<br />

From AAAS 2010<br />

NEWS OF THE WEEK<br />

Science reporters posted more than two<br />

dozen blog entries and podcasts from the<br />

meeting. Here is a sample. For full coverage,<br />

see www.sciencenow.org. And to see what<br />

our guest bloggers had to say, see<br />

news.sciencemag.org/sciencebloggers.<br />

A Sexy Treatment for Traumatic Brain Injury<br />

The hormone progesterone is best known for<br />

its work in the female reproductive system,<br />

where it plays various roles in supporting<br />

pregnancy. But starting next month, it will be<br />

the focus of a phase III clinical trial for traumatic<br />

brain injury. Researchers hope an infusion<br />

of progesterone given within a few hours<br />

of a car accident or other trauma will help<br />

prevent brain damage.<br />

The Mathematics of Clumpy Crime<br />

Even in a sprawling city like Los Angeles,<br />

California, crimes still clump together. Mathematical<br />

models presented at the meeting<br />

show that such crime hot spots form when<br />

previous crimes attract more criminals to a<br />

neighborhood. By understanding how these<br />

blobs form, researchers hope to help police<br />

break them up.<br />

Are ‘Test Tube Babies’ Healthy?<br />

When Louise Brown was born on 25 July<br />

1978, she kicked off an era. The first “test<br />

tube baby” is a mother herself now, and she’s<br />

been joined by millions of others born with<br />

the help of in vitro fertilization (IVF). But are<br />

babies born via IVF the same as those born<br />

naturally? Researchers have discovered some<br />

subtle genetic differences.<br />

Drive Green, Make Money<br />

Widespread adoption of plug-in electric vehicles<br />

could dramatically cut greenhouse gas<br />

pollution and reduce U.S. dependence on<br />

foreign oil. And results of a new electric-car<br />

pilot project provide added incentive to go<br />

electric: Car owners could return unused<br />

electricity to the grid and make real money<br />

doing so.<br />

Science Is Kryptonite for Superheroes<br />

Hollywood has a message for scientists: If<br />

you want something that’s 100% accurate,<br />

go watch a documentary. A panel of screenwriters<br />

for superhero-driven movies and TV<br />

shows like Watchmen and Heroes said that<br />

their job is to get the characters right, not<br />

the science.<br />

www.sciencemag.org SCIENCE VOL 327 26 FEBRUARY 2010 1071<br />

Published by AAAS<br />

on March 11, 2010<br />

www.sciencemag.org<br />

Downloaded from


Functional Ecology 2008, 22,<br />

284<strong>–</strong>288 doi: 10.1111/j.1365-2435.2007.01354.x<br />

Blackwell Publishing Ltd<br />

Infant carrying behaviour in dolphins: costly parental<br />

care in an aquatic environment<br />

S. R. Noren*<br />

Institute of Marine Science, Center for Ocean Health, University of California at Santa Cruz, 100 Shaffer Road, Santa Cruz,<br />

CA 95060, USA<br />

Functional doi: 10.1111/j.1365-2435.2007.0@@@@.x<br />

Ecology (2007)<br />

xx,<br />

000<strong>–</strong>000<br />

Summary<br />

1. Infant carrying behaviour occurs across diverse taxa inhabiting arboreal, volant and aquatic<br />

environments. For mammals, it is considered to be the most expensive form of parental care after<br />

lactation, yet the effect of infant carrying on the energetics and performance of the carrier is<br />

virtually unknown.<br />

2. Echelon swimming in cetacean (dolphin and whale) mother<strong>–</strong>infant dyads, described as calf in<br />

very close proximity of its mother’s mid-lateral flank, appears to be a form of aquatic ‘infant<br />

carrying’ behaviour as indicated by the hydrodynamic benefits gained by calves in this position<br />

which enables them to maintain proximity of their travelling mothers. Although this behaviour<br />

provides a solution for minimizing separations of mother<strong>–</strong>infant dyads, it may be associated with<br />

maternal costs.<br />

3. Through kinematic analyses this study demonstrates empirically that ‘infant carrying’ impacts<br />

the locomotion of dolphin ( Tursiops truncatus)<br />

mothers as evident by decreased swim performance<br />

and increased effort.<br />

4. The mean maximum swim speed of mothers swimming in echelon only represented 76% of the<br />

mean maximum swim speed of these mothers swimming solitarily. In addition, there was a<br />

concomitant 13% reduction in distance per stroke for mothers swimming in echelon compared to<br />

periods of solitary swimming.<br />

5. Thus, ‘infant carrying’ in an aquatic environment is associated with maternal costs, and could<br />

ultimately impact maternal energy budgets, foraging efficiency and predator evasion.<br />

Introduction<br />

Key-words:<br />

cetacean, echelon position, hydrodynamics, kinematics, swimming<br />

Maternal investment in mammals includes gestation,<br />

lactation and other forms of parental care. Infant carrying is<br />

considered to be the most costly form of parental care after<br />

lactation (Altmann & Samuels 1992; Kramer 1998) and has<br />

been described in 6 of 19 eutherian mammalian orders (for<br />

review, see Ross 2001). This behaviour provides a solution for<br />

mothers of diverse taxa that must manoeuvre within their<br />

environment to forage and avoid predators while accompanied<br />

by their young offspring (Ross 2001), which are<br />

handicapped by small body size, undeveloped tissues and<br />

naïveté (Carrier 1996). This behaviour is only thought to<br />

evolve when offspring are unable to independently follow<br />

their mothers, as in arboreal (i.e. primates) and volant (i.e.<br />

bats) environments (Ross 2001).<br />

*Correspondence author. E-mail: snoren@biology.ucsc.edu<br />

The aquatic environment also appears to require ‘infant<br />

carrying’ to ensure that mother<strong>–</strong>infant dyads remain intact<br />

during travel as manatees and sea otters are observed<br />

physically carrying their young. Cetaceans (whale and dolphin),<br />

however, cannot physically carry their young. Yet similar to<br />

that observed for primates (Altmann & Samuels 1992; Doran<br />

1992; Wells & Turnquist 2001), mature locomotor performance<br />

in dolphins is precluded for several years postpartum (Noren,<br />

Biedenbach & Edwards 2006). Echelon position is the<br />

predominant behaviour displayed by cetacean mother<strong>–</strong>infant<br />

dyads (Fig. 1; McBride & Kritzler 1951; Tavolga & Essapian<br />

1957; Norris & Prescott 1961; Au & Perryman 1982; Taber &<br />

Thomas 1982; Mann & Smuts 1999; Noren & Edwards 2007)<br />

and appears to represent an aquatic form of ‘infant carrying’<br />

because it enables neonatal cetaceans to maintain close<br />

proximity to their mothers during travel (Norris & Prescott<br />

1961; Lang 1966) by increasing the swimming efficiency of the<br />

infant (Kelly 1959; Weihs 2004; Noren et al.<br />

2008). Thus<br />

cetaceans, like primates, appear to ‘carry’ their young.<br />

© 2007 The Author. Journal compilation © 2007 British Ecological Society


Fig. 1. Three bottlenose dolphin mother<strong>–</strong>calf pairs swimming in echelon<br />

position. Echelon position is described as calf in very close proximity<br />

with its mother’s mid-lateral flank in the region near her dorsal fin.<br />

Photo © and courtesy of Dolphin Quest Hawaii.<br />

Only a few studies have examined the energetic and<br />

locomotor consequences of infant carrying for the carrier,<br />

and these studies have focused on primates (Altmann &<br />

Samuels 1992; Schradin & Anzenberger 2001) and marsupials<br />

(Baudinette & Biewener 1998). Meanwhile the maternal<br />

consequences of infant carrying in other taxa and environments<br />

(i.e. volant and aquatic) remain unexplored. In view of this, I<br />

examined the kinematics of dolphin mothers swimming with<br />

their calf in echelon (Fig. 1) and swimming solitarily (> 1 m<br />

from their calf and all other dolphins) to elucidate the effect of<br />

‘infant carrying’ on maternal locomotor performance and<br />

effort in an aquatic environment. The advantages gained by<br />

calves in echelon position are examined in the accompanying<br />

paper (Noren et al.<br />

2008).<br />

Materials and methods<br />

Three captive bottlenose dolphin ( Tursiops truncatus)<br />

mother<strong>–</strong>calf<br />

pairs housed at Dolphin Quest Hawaii provided a controlled<br />

experimental approach to investigate cetacean locomotor performance<br />

and effort (Fish 1993; Skrovan et al.<br />

1999; Noren et al.<br />

2006, 2008).<br />

Methodologies regarding the dolphin enclosure, placement of the<br />

SCUBA diver-videographer, and placement of the dolphins in the<br />

water column are described in detail elsewhere (Noren et al.<br />

2008).<br />

Experimental swim sessions included both opportunistic (no reward)<br />

and directional swimming between two trainers (reward-based).<br />

Echelon swimming was recorded only when the mothers’ calves<br />

were 0<strong>–</strong>34 days postpartum. Thus, the present study only provides<br />

details regarding the maternal costs of echelon for mothers with<br />

very young calves (0<strong>–</strong>34 days postpartum). Solitary swimming was<br />

recorded at several intervals 0<strong>–</strong>2 years past parturition. Thirty-three<br />

hours of swimming were recorded and 334 short 1<strong>–</strong>6 s video clips<br />

were extracted and digitized (Fig. 2). Clips were divided into two<br />

association categories: (i) echelon position (Fig. 1); and (ii) solitary<br />

swimming (mother > 1 m away from calf and all other dolphins). In<br />

all clips the mothers were continuously stroking.<br />

A quantitative assessment of swim effort was obtained by<br />

calculating peak-to-peak fluke stroke amplitude and tailbeat<br />

oscillation frequency. Higher amplitudes and frequencies are<br />

associated with greater energy expenditure (Kooyman & Ponganis<br />

1998). Normalized tailbeat frequency (ratio of tailbeat frequency to<br />

© 2007 The Author. Journal compilation © 2007 British Ecological Society, Functional Ecology,<br />

22,<br />

284<strong>–</strong>288<br />

Infant carrying in dolphins<br />

285<br />

Fig. 2. A tracing from a digitized video clip of a solitarily swimming<br />

mother dolphin. Anatomical points of interest (rostrum tip, cranial<br />

insertion of the dorsal fin, and fluke tip) were digitized at a rate of 60<br />

fields per second of video using a motion-analysis system (Peak<br />

Motus 6·1; Peak Performance Technologies, Inc. Englewood, CO,<br />

USA) following methods similar to Skrovan et al. (1999) and Noren<br />

et al. (2006). A distinct trace represents the movements of each<br />

digitised anatomical point. From left to right, the trace from the<br />

rostrum leads (pink), followed by the trace from the cranial insertion<br />

of the dorsal fin (yellow), and last is the trace from the fluke tip (blue).<br />

The brown dot is a digitized reference point indicating that the<br />

camera was steady while filming this video clip.<br />

swim speed; Rohr & Fish 2004) and distance per stroke were also<br />

calculated. Methods for video analysis and swim effort calculations<br />

are described in detail elsewhere (Noren et al.<br />

2006).<br />

The goal of this study was not to address individual variation, but<br />

to quantify changes in locomotor performance associated with swim<br />

style (echelon vs. solitary), thus similar to other kinematic studies<br />

data across individuals were pooled (Fish 1993; Skrovan et al.<br />

1999;<br />

Noren et al.<br />

2006, 2008). Pearson product moment correlation<br />

coefficients were used to determine the correlations of peak-to-peak<br />

fluke stroke amplitude and tailbeat frequency with swim speed for<br />

echelon swimming and also for solitary swimming; linear regression<br />

analyses were then used to determine the relationship for the parameters<br />

that demonstrated a strong correlation. Swim speed, normalized<br />

tailbeat frequency, and distance per stroke during echelon swimming<br />

and solitary swimming were compared using student’s t-tests<br />

when normally distributed, or Mann<strong>–</strong>Whitney rank sum tests when<br />

normality failed ( α = 0·05). The maximum swim performance of<br />

each individual during echelon swimming and solitary swimming<br />

was compared using a paired t-test<br />

( α = 0·10). Statistical analyses<br />

were performed using Sigma<br />

Stat<br />

2·03 (Systat Software, Inc. Point<br />

Richmond, CA, USA). Means ± 1 SEM are presented.<br />

Results<br />

Our experimental approach adequately captured swimming<br />

behaviours representative of wild dolphins (Noren et al.<br />

2008). The average swim speed of mother dolphins was<br />

<strong>–</strong>1<br />

significantly slower during echelon swimming (2·11 ± 0·06 m s ;<br />

<strong>–</strong>1<br />

n = <strong>17</strong>8) than during solitary swimming (3·88 ± 0·09 m s ;<br />

n = 156; T = 36534·00, P < 0·001). Given that 53% of the<br />

echelon swim data were from directional swim trials, compared<br />

to 94% for the solitary swim data, swim speed data from<br />

directional trials only were also compared to ensure that the<br />

previous result was not due to experimental design. The result


286<br />

S. R. Noren<br />

Fig. 3. Swimming kinematics of the mother in relation to the<br />

swimming speed of the mother. Peak-to-peak fluke stroke amplitude<br />

(a) and tailbeat frequency (b) were both correlated with swim speed<br />

for mother dolphins swimming in echelon with their calf (white<br />

symbols; r = 0·400, P < 0·001, n = <strong>17</strong>8 and r = 0·799, P < 0·001, n =<br />

<strong>17</strong>8, respectively) and swimming solitarily (black symbols; r = 0·277,<br />

P < 0·001, n = 156 and r = 0·885, P < 0·001, n = 156, respectively).<br />

The relationship between swim speed and peak-to-peak fluke stroke<br />

amplitude appears to be nonlinear. Given the strong linear correlation<br />

between swim speed and tailbeat frequency (SF) linear regressions<br />

are provided for echelon (speed = 1·61 SF + 0·27; r 2 = 0·638, F =<br />

310·794, P < 0·001) and solitary (speed = 2·09 SF + 0·13; r 2 = 0·783,<br />

F = 555·610, P < 0·001) swimming. A different symbol is used for<br />

each of the three individual dolphins.<br />

was the same; average swim speed during echelon swimming<br />

was significantly slower than during solitary swimming<br />

( t = −11·644,<br />

P < 0·001, n = 94, 145). The absolute maximum<br />

swim speed for each mother swimming with their calf in<br />

<strong>–</strong>1<br />

echelon was 4·23, 4·39 and 4·32 m s for animals 1, 2 and 3,<br />

<strong>–</strong>1<br />

respectively. This compares to 5·65, 6·32 and 5·11 m s for<br />

animals 1, 2 and 3 swimming solitarily, respectively. As a<br />

result, mean maximum swim performance was significantly<br />

<strong>–</strong>1<br />

slower when mothers were swimming in echelon (4·31 ± 0·05 m s )<br />

<strong>–</strong>1<br />

compared to solitary swim periods (5·69 ± 0·35 m s ; t = −4·816,<br />

n = 3, P = 0·053).<br />

Peak-to-peak fluke stroke amplitudes (Fig. 3a) and tailbeat<br />

frequencies (Fig. 3b) were correlated with swim speed for<br />

mother dolphins swimming in echelon and swimming<br />

solitarily. Because average swim speed was significantly different<br />

between the two swim categories, swim effort was standardized<br />

for swim speed. Mothers in echelon demonstrated significant<br />

increases in effort compared to periods of solitary swimming<br />

as evident by greater normalized tailbeat frequency<br />

Fig. 4. ‘Infant carrying’ imparted maternal costs. Mother dolphins<br />

swimming in echelon with their calf (white bar) demonstrated<br />

significantly lower mean distance covered per stroke (t = −7·068,<br />

P < 0·001, n = <strong>17</strong>8, 156) compared to periods of solitary swimming<br />

(black bars).<br />

( T = 19949·00, P < 0·001, n = <strong>17</strong>8, 156) and reduced distance<br />

per stroke (Fig. 4). For a given speed, mothers swimming in<br />

echelon increased tailbeat frequency by <strong>17</strong>%, which resulted<br />

in a 13% decrease in distance covered per stroke compared to<br />

solitary swimming.<br />

Discussion<br />

Infant carrying provides a solution for mothers who must<br />

locomote in their environment while accompanied by their<br />

young offspring. Echelon swimming in cetacean mother<strong>–</strong><br />

infant dyads is a type of ‘infant carrying’ that improves calf<br />

swim performance (Noren et al.<br />

2008), but it appears to come<br />

with a maternal cost. Average and maximum swim speeds<br />

were significantly slower for mothers swimming in echelon<br />

with their calves compared to periods of solitary swimming.<br />

For example, the mean maximum swim speed of the mothers<br />

swimming in echelon only represented 76% of the mean<br />

maximum performance of these mothers swimming solitarily.<br />

The maximum speed was assumed to represent the animal’s<br />

extreme performance, which is the method used to qualify<br />

physiological capacity (Weibel et al.<br />

1987), thus this result<br />

implies that the presence of a calf is a detriment to maternal<br />

swim performance.<br />

An alternate hypothesis for the decreased performance of<br />

dolphin mothers swimming in echelon is that the mother can<br />

only travel as fast as the infant can actively swim because the<br />

infant must occasionally stroke during the echelon swim<br />

behaviour. However, 0<strong>–</strong>1 month-old calves are capable of<br />

<strong>–</strong>1<br />

independent swim speeds of 0·58<strong>–</strong>4·20 m s (Noren et al.<br />

2006) and this performance is increased when the calf is in<br />

echelon (Noren et al.<br />

2008) because it receives up to 60% of<br />

the thrust from its mother (Weihs 2004). Thus, it is unlikely<br />

that the swimming ability of an entrained calf constrains<br />

maternal echelon swim performance. Studies of chimpanzees<br />

( Pan troglodytes)<br />

also suggest that infant carrying decreases<br />

maternal travel speed (Wrangham 2000; Williams, Hsien-Yang<br />

& Pusey 2002).<br />

© 2007 The Author. Journal compilation © 2007 British Ecological Society, Functional Ecology,<br />

22,<br />

284<strong>–</strong>288


In addition to decreased performance, dolphin mothers<br />

swimming in echelon were required to increase effort compared<br />

to periods of solitary swimming. For a given speed,<br />

mothers swimming in echelon significantly increased tailbeat<br />

frequency with the result that distance covered per stroke<br />

significantly decreased by 13% compared to solitary<br />

swimming (Fig. 4). Interestingly, the proportion of decreased<br />

distance per stroke in ‘infant carrying’ dolphin mothers is<br />

strikingly similar to the <strong>17</strong>% decrease in distance per leap<br />

measured empirically for marmosets ( Callithrix jacchus)<br />

carrying weights equivalent to their newborn twin offspring<br />

(Schradin & Anzenberger 2001). Furthermore, although<br />

female wallabies carrying a load (approximating the mass of<br />

a fully developed offspring) did not alter stride frequency,<br />

they increased the time the foot applied force on the ground,<br />

which likely increased the stored elastic strain energy to levels<br />

necessary to transport the additional load (Baudinette &<br />

Biewener 1998). Regardless of habitat (aquatic vs. arboreal)<br />

or forces to overcome (hydrodynamic drag vs. gravity) in these<br />

systems maternal effort must change to support an increased load.<br />

The increase in maternal effort for dolphin mothers<br />

swimming in echelon compared to periods of solitary swimming<br />

may be associated with changes in water flow patterns and<br />

drag. The presence of the calf may disrupt the boundary flow<br />

around the mother causing it to separate, which would<br />

increase turbulent flow. In addition, the entrained calf could<br />

increase the surface area of the mother, which effectively<br />

increases the drag of the swimmer (Webb 1975). More power<br />

is required to overcome increased turbulent flow and drag<br />

(Webb 1975). As a greater proportion of maternal power<br />

output is utilized to accommodate increased turbulent flow<br />

and drag, there is less energy available to propel the animal<br />

forward. As a result, locomotor performance decreases<br />

because power output per stroke is limited by mechanical<br />

constraints (Fish & Hui 1991) and total work is limited by the<br />

animal’s metabolic scope (Weibel et al.<br />

1987). These relationships<br />

may explain the observed decrease in swim speed and<br />

distance covered per stroke for mothers swimming in echelon<br />

position. Although Weihs (2006) suggested that the gain in<br />

forward forces by the following body (calf) is larger than the<br />

added cost to the leading body (mother), a more detailed<br />

theoretical examination of the drag and flow patterns for the<br />

leading body in echelon position is warranted to validate the<br />

hypotheses for the proposed decreased maternal performance<br />

observed in the present study.<br />

Given that infant carrying mothers must forage and evade<br />

predators, one of many constraints on newborn offspring<br />

mass may be its impact on maternal locomotor performance.<br />

As a result, newborn dolphins and marmosets represent a<br />

similar proportion of maternal mass, 15% (mass data from a<br />

mother<strong>–</strong>calf pair in this study) and <strong>17</strong>% (Tardif, Harrison &<br />

Simek 1993), respectively. Ultimately as an infant ages and<br />

increases in size, maternal costs associated with infant<br />

carrying theoretically increase in aquatic (Weihs 2004, 2006)<br />

and arboreal (Schradin & Anzenberger 2001) environments.<br />

Optimal theory predicts that maternal carrying costs should<br />

not outweigh the sum of the costs of independent locomotion<br />

© 2007 The Author. Journal compilation © 2007 British Ecological Society, Functional Ecology,<br />

22,<br />

284<strong>–</strong>288<br />

Infant carrying in dolphins<br />

287<br />

by the mother and her offspring (Kramer 1998). Therefore,<br />

the transition to offspring locomotor independence may<br />

represent a parent<strong>–</strong>offspring conflict (Trivers 1974) as energy<br />

expenditure to carry an infant may limit the mother’s available<br />

energy to invest in future offspring (Kramer 1998). As<br />

cetacean and primate offspring increase in size, there is an<br />

increase in the active prevention and/or avoidance of infant<br />

carrying by mothers in both groups (Altmann 1980; Taber &<br />

Thomas 1982; Mann & Smuts 1999), such that with age there<br />

is a decrease in the time cetacean infants swim in echelon<br />

(Taber & Thomas 1982; Mann & Smuts 1999; Noren &<br />

Edwards 2007) and primate infants are carried (Altmann &<br />

Samuels 1992; Salvage et al.<br />

1996; Pontzer & Wrangham<br />

2006). Examination of human infant carrying behaviour also<br />

suggests that by a certain mass and age, mothers encourage<br />

their infants to walk independently (Kramer 1998).<br />

In summary, this study provides the first empirical evidence<br />

of the maternal consequences of ‘infant carrying’ in an aquatic<br />

environment. Although infant carrying provides a solution<br />

for mothers that must manoeuvre within their environment<br />

while accompanied by their underdeveloped offspring, this<br />

behaviour is associated with maternal costs regardless of environment,<br />

as evident in arboreal (Schradin & Anzenberger 2001)<br />

and aquatic (present study) regimes. The decreased locomotor<br />

performance and increased locomotor effort associated with<br />

infant carrying undoubtedly impacts maternal energy budgets,<br />

foraging efficiency and predator evasion. Given the prevalence<br />

of infant carrying behaviour across diverse taxa and habitats<br />

and the consequences of this behaviour, it is surprising that<br />

the energetics of infant carrying have largely been ignored.<br />

Future investigations are warranted, particularly in a volant<br />

environment, which remains unexplored.<br />

Acknowledgments<br />

I thank Dolphin Quest, particularly J. Sweeney and R. Stone, for providing the<br />

experimental facilities and animals and for funding portions of data collection<br />

and analyses. I also thank Southwest Fisheries Science Center (SWFSC),<br />

particularly S. Reilly and E. Edwards, for funding portions of data collection.<br />

In addition, I thank the staff at Dolphin Quest Hawaii (particularly G. Biedenbach<br />

and C. Buczyna), T. Williams of the University of California Santa Cruz for the<br />

use of her Peak Motus system, and J. Redfern of SWFSC for assistance with<br />

data management.<br />

References<br />

Altmann, J. (1980) Baboon Mothers and Infants.<br />

Harvard University Press,<br />

Cambridge.<br />

Altmann, J. & Samuels, A. (1992) Costs of maternal care: infant-carrying<br />

baboons. Behavioural Ecology and Sociobiology,<br />

29,<br />

391<strong>–</strong>398.<br />

Au, D. & Perryman, W. (1982) Movement and speed of dolphin schools<br />

responding to an approaching ship. Fishery Bulletin,<br />

80,<br />

371<strong>–</strong>379.<br />

Baudinette, R.V. & Biewener, A.A. (1998) Young wallabies get a free ride.<br />

Nature,<br />

395,<br />

653<strong>–</strong>654.<br />

Carrier, D.R. (1996) Ontogenetic limits on locomotor performance. Physiological<br />

Zoology,<br />

69,<br />

467<strong>–</strong>488.<br />

Doran, D.M. (1992) The ontogeny of chimpanzee and pygmy chimpanzee<br />

locomotor behavior: a case study of paedomorphism and its behavorial<br />

correlates. Journal of Human Evolution,<br />

23,<br />

139<strong>–</strong>158.<br />

Fish, F.E. (1993) Power output and propulsive efficiency of swimming bottlenose<br />

dolphins ( Tursiops truncatus).<br />

Journal of Experimental Biology,<br />

185,<br />

<strong>17</strong>9<strong>–</strong>193.<br />

Fish, F.E. & Hui, C.A. (1991) Dolphin swimming <strong>–</strong> a review. Mammal Review,<br />

21,<br />

181<strong>–</strong>195.


288<br />

S. R. Noren<br />

Kelly, H.R. (1959) A two body problem in the echelon swimming of porpoise.<br />

Naval Ordinance Test Station Technical Note 40606<strong>–</strong>1.<br />

Kooyman, G.L. & Ponganis, P.J. (1998) The physiological basis for diving at<br />

depth: birds and mammals. Annual Review of Physiology 60,<br />

19<strong>–</strong>32.<br />

Kramer, P.A. (1998) The costs of human locomotion: maternal investment in<br />

child transport. American Journal of Physical Anthropology,<br />

107,<br />

71<strong>–</strong>85.<br />

Lang, T.G. (1966) Hydrodynamic analysis of cetacean performance. Whales,<br />

Dolphins, and Porpoises (ed. K.S. Norris), pp. 410<strong>–</strong>432. University of<br />

California Press, Berkeley.<br />

Mann, J. & Smuts, B. (1999) Behavioral development in wild bottlenose<br />

dolphin newborns ( Tursiops sp.). Behavior,<br />

136,<br />

529<strong>–</strong>566.<br />

McBride, A.F. & Kritzler, H. (1951) Observations on pregnancy, parturition,<br />

and post-natal behavior in the bottlenose dolphin. Journal of Mammology,<br />

32,<br />

251<strong>–</strong>266.<br />

Noren, S.R. & Edwards, E.F. (2007) Physiological and behavioral development<br />

in dolphin calves: implications for calf separation and mortality due to tuna<br />

purse-seine sets. Marine Mammal Science,<br />

23(1),<br />

15<strong>–</strong>29.<br />

Noren, S.R., Biedenbach, G. & Edwards, E.F. (2006) The ontogeny of swim<br />

performance and mechanics in bottlenose dolphins ( Tursiops truncatus).<br />

Journal of Experimental Biology,<br />

209(23),<br />

4724<strong>–</strong>4731.<br />

Noren, S.R., Biedenbach, G., Redfern, J.V. & Edwards, E.F. (2008) Hitching a<br />

ride: the formation locomotion strategy of dolphin calves. Functional<br />

Ecology.<br />

doi: 10.1111/j.1365-2435.2007.01353.x.<br />

Norris, K.S. & Prescott, J.H. (1961) Observations on Pacific cetaceans of<br />

Californian and Mexican waters. University of California Publications in<br />

Zoology, 63, 291<strong>–</strong>402.<br />

Pontzer, H. & Wrangham, R.W. (2006) Ontogeny of ranging in wild chimpanzees.<br />

<strong>International</strong> Journal of Primatology, 27(1), 295<strong>–</strong>309.<br />

Rohr, J.R. & Fish, F.E. (2004) Strouhal numbers and optimization of swimming<br />

by odontocete cetaceans. Journal of Experimental Biology, 207, 1633<strong>–</strong>1642.<br />

Ross, C. (2001) Park or ride? Evolution of infant carrying in primates. <strong>International</strong><br />

Journal of Primatology, 22(5), 749<strong>–</strong>771.<br />

Salvage, A., Snowden, P.T., Giraldo, L.H. & Soto, L.H. (1996) Parental care<br />

patterns and vigilance in wild cotton-top tamarins (Saguinus oedipus).<br />

Adaptive Radiation of Neotropical Primates (eds N.A. Norconk, A.L.<br />

Rosenberg & P.A. Garber), pp. 187<strong>–</strong>199. Plenum Press, New York.<br />

Schradin, C. & Anzenberger, G. (2001) Costs of infant carrying in common<br />

marmosets, Callithrix jacchus: an experimental analysis. Animal Behavior,<br />

62, 289<strong>–</strong>295.<br />

Skrovan, R.C., Williams, T.M., Berry, P.S. & Moore, P.W. (1999) The diving<br />

physiology of bottlenose dolphins (Tursiops truncatus) II. Biomechanics and<br />

changes in buoyancy at depth. Journal of Experimental Biology, 202, 2749<strong>–</strong><br />

2761.<br />

Taber, S. & Thomas, P. (1982) Calf development and mother<strong>–</strong>calf spatial<br />

relationships in southern right whales. Animal Behaviour, 30, 1072<strong>–</strong>1083.<br />

Tardif, S.D., Harrison, M.L. & Simek, M.A. (1993) Communal infant care in<br />

marmosets and tamarins: relation to energetics, ecology, and social organization.<br />

Marmosets and Tamarins: Systematics, Behavior, and Ecology (ed.<br />

A.B. Rylands), pp. 200<strong>–</strong>219. Oxford Press, Oxford.<br />

Tavolga, M.C. & Essapian, F.S. (1957) The behavior of the bottle-nosed<br />

dolphin (Tursiops truncatus): mating, pregnancy, parturition and mother<strong>–</strong><br />

infant behavior. Zoologica, 42, 11<strong>–</strong>31.<br />

Trivers, R. (1974) Parent-offspring conflict. American Zoologist, 14, 249<strong>–</strong>264.<br />

Webb, P.W. (1975) Hydrodynamics and energetics of fish propulsion. Bulletin<br />

of the Fisheries Research Board of Canada, 190, 1<strong>–</strong>159.<br />

Weibel, E.R., Taylor, C.R., Hoppeler, H. & Karas, R.H. (1987) Adaptive<br />

variation in the mammalian respiratory system in relation to energetic<br />

demand: I. Introduction to problem and strategy. Respiration Physiology,<br />

69, 1<strong>–</strong>6.<br />

Weihs, D. (2004) The hydrodynamics of dolphin drafting. Journal of Biology,<br />

3(8), 1<strong>–</strong>23 (http://jbiol.com/content/3/2/8).<br />

Weihs, D. (2006) Aerodynamic interactions between adjacent slender bodies.<br />

American Institute of Aeronautics and Astronautics Journal, 44, 481<strong>–</strong>484.<br />

Wells, J.P. & Turnquist, J.E. (2001) Ontogeny of locomotion in rhesus macaques<br />

(Macaca mulatta): II postural and locomotor behavior and habitat use in a<br />

free-ranging colony. American Journal of Physical Anthropology, 115, 80<strong>–</strong><br />

94.<br />

Williams, J.M., Hsien-Yang, L. & Pusey, A.E. (2002) Costs and benefits of<br />

grouping for female chimpanzees at Gombe. Behavioural Diversity in<br />

Chimpanzees and Bonobos (ed. C. Boesch), pp. 192<strong>–</strong>203. Cambridge University<br />

Press, Cambridge.<br />

Wrangham, R.W. (2000) Why are male chimpanzees more gregarious than<br />

mothers? A scramble competition hypothesis. Primate Males, Causes and<br />

Consequences of Variation in Group Composition (ed. P.M. Kaeppler),<br />

pp. 248<strong>–</strong>258. Cambridge University Press, Cambridge.<br />

Received 12 June 2007; accepted 27 September 2007<br />

Handling Editor: Francisco Bozinovic<br />

© 2007 The Author. Journal compilation © 2007 British Ecological Society, Functional Ecology,<br />

22,<br />

284<strong>–</strong>288


MARINE MAMMAL SCIENCE, 23(3): 629<strong>–</strong>649 (July 2007)<br />

C○ 2007 by the Society for Marine Mammalogy<br />

DOI: 10.1111/j.<strong>17</strong>48-7692.2007.00133.x<br />

SIMULATED VESSEL APPROACHES ELICIT<br />

DIFFERENTIAL RESPONSES FROM MANATEES<br />

JENNIFER L. MIKSIS-OLDS 1<br />

PERCY L. DONAGHAY<br />

Graduate School of Oceanography,<br />

University of Rhode Island,<br />

Narragansett, Rhode Island 02882, U.S.A.<br />

E-mail: jmiksis@gso.uri.edu<br />

JAMES H. MILLER<br />

Department of Ocean Engineering<br />

and<br />

Graduate School of Oceanography,<br />

University of Rhode Island,<br />

Narragansett, Rhode Island 02882, U.S.A.<br />

PETER L. TYACK<br />

Woods Hole Oceanographic Institution,<br />

Woods Hole, Massachusetts 02543, U.S.A.<br />

JOHN E. REYNOLDS, III<br />

Mote Marine Laboratory,<br />

1600 Ken Thompson Parkway,<br />

Sarasota, Florida 34236, U.S.A.<br />

ABSTRACT<br />

One of the most pressing concerns associated with conservation of the endangered<br />

Florida manatee is mortality and serious injury due to collisions with watercraft.<br />

Watercraft collisions are the leading identified cause of manatee mortality, averaging<br />

25% and reaching 31% of deaths each year. The successful establishment and<br />

management of protected areas depend upon the acquisition of data assessing how<br />

manatees use different habitats, and identification of environmental characteristics<br />

influencing manatee behavior and habitat selection. Acoustic playback experiments<br />

were conducted to assess the behavioral responses of manatees to watercraft approaches.<br />

Playback stimuli made from prerecorded watercraft approaches were constructed<br />

to simulate a vessel approach to approximately 10 m in sea grass habitats.<br />

Stimulus categories were (1) silent control, (2) approach with outboard at idle speed,<br />

(3) vessel approach at planning speed, and (4) fast personal watercraft approach.<br />

Analyses of swim speed, changes in behavioral state, and respiration rate indicate<br />

1 Present address of corresponding author: Jennifer L. Miksis-Olds, The Pennsylvania State University,<br />

Applied Research Laboratory, P. O. Box 30, State College, Pennsylvania 16804, U.S.A.<br />

629


630 MARINE MAMMAL SCIENCE, VOL. 23, NO. 3, 2007<br />

that the animals responded differentially to the playback categories. The most pronounced<br />

responses, relative to the controls, were elicited by personal watercraft.<br />

Quantitative documentation of response during playbacks provides data that may<br />

be used as the basis for future models to predict the impact of specific human<br />

activities on manatees and other marine mammal populations.<br />

Key words: manatee, Trichechus manatus latirostris, disturbance, avoidance, playback<br />

experiment, watercraft approaches.<br />

A pressing question regarding the conservation of Florida manatees (Trichechus<br />

manatus latirostris) ishow to minimize both the lethal and nonlethal impacts on<br />

manatees in areas where watercraft operate. The lethal impacts of watercraft account<br />

for an average of 25% to more than 30% of identified manatee deaths each year<br />

(Ackerman et al. 1995, Reynolds 1999). The nonlethal impacts of boating on the<br />

physical health of manatees, as well as an indirect impact on food availability and<br />

communication, are controversial topics for which available data are inadequate.<br />

Identifying specific environments or behaviors that put manatees at a greater risk<br />

for boat collisions and quantifying manatee reactions to varying speeds and motor<br />

types of approaching vessels are necessary steps toward achieving the overall goal<br />

of minimizing the negative, lethal, and nonlethal impacts of boats and associated<br />

noise.<br />

Manatees most commonly encounter relatively small boats: outboard or inboard/outboard<br />

leisure boats, personal watercraft (PWC), and fishing trawlers<br />

(Gorzelany 2004). PWC means a vessel


MIKSIS-OLDS ET AL.: SIMULATED VESSEL APPROACHES 631<br />

Specifically, swimming speed increased and animals moved toward deeper channel<br />

waters in response to boat approaches. The distance from the manatees to the boat,<br />

water depth at the boat, and water depth at the manatees each had a significant effect<br />

on swimming speed, whereas the type of boat or boat speed did not have a significant<br />

effect.<br />

Accurate detection of an approaching boat is only part of the problem for a manatee.<br />

In order to swim out of the direct path of the boat, the manatee must accurately<br />

localize the boat and respond appropriately. The physical characteristics of sound<br />

propagation in shallow water are complicated, and this can make it difficult to locate<br />

a sound source. From an acoustic standpoint, shallow water refers to areas where sound<br />

is propagated to distances at least several times the water depth, under conditions<br />

where both the surface and sediment boundaries have an effect on transmission (Urick<br />

1983). Compared to deep-water environments, there is greater attenuation of sound<br />

in shallow water environments (Medwin and Clay 1998). Higher frequencies have<br />

shorter wavelengths and are therefore more directional than lower frequencies with<br />

respect to a sound source of a given size. High levels of sound reverberation in shallow<br />

water also make localization difficult for even those species that possess extremely<br />

accurate localization ability. For example, bottlenose dolphins (Tursiops truncatus),<br />

which have excellent acoustic localization capabilities, have been hit by boats in shallow<br />

water (Wells and Scott 1997, Buckstaff 2004). These observations suggest that<br />

successful localization of approaching boats is necessary, but not a sufficient component,<br />

for dolphins to safely avoid approaching watercraft. Localizing approaching<br />

boats may also be an important component for manatees to avoid collision.<br />

Manatees that do manage to avoid collisions with watercraft still need to cope<br />

with other effects of boats. Boats that produce noise over the same frequencies as<br />

manatee vocalizations (Nowacek et al. 2003) can potentially mask communication<br />

signals. Motor noise can potentially mask communication signals. For example, if<br />

boat noise interferes with communication, females may lose contact with their calves<br />

or other members of a group, which could affect survival of the calves (Bengtson and<br />

Fitzgerald 1985). Manatees also are displaced from critical habitats with chronic boat<br />

disturbance (Provancha and Provancha 1988, Buckingham et al. 1999). Heavy vessel<br />

traffic may also cause manatees to expend more energy than they normally would<br />

(Reynolds 1999). For example, manatees significantly increase swim speed and move<br />

from shallow habitats to deeper channels during boat approaches (Nowacek et al.<br />

2001a). Based on these findings, as well as observations of behavioral responses to<br />

opportunistic vessel approaches prior to the initiation of this study, three categories of<br />

response (slow swim, fast swim, and rolling dive) were defined prior to the initiation<br />

of the study. The magnitude and diversity of manatee responses to vessel traffic clearly<br />

indicate the need for playback experiments to demonstrate manatee responses to the<br />

acoustic component of vessel approaches.<br />

Stimulus Recordings<br />

MATERIALS AND METHODS<br />

All stimulus recordings were made using a single hydrophone suspended from<br />

a recording vessel anchored in a sea grass habitat adjacent to a boating channel.<br />

The recording hydrophone was a HTI-99-HF hydrophone with built-in preamplifier<br />

and had a 2 Hz<strong>–</strong>125 kHz frequency range and −<strong>17</strong>8 dB re 1V/Pa sensitivity.<br />

The recording system was a National Instruments PCMCIA DAQ Card-6062E used


632 MARINE MAMMAL SCIENCE, VOL. 23, NO. 3, 2007<br />

Table 1. Characteristics of playback stimuli<br />

Category Exemplar<br />

Received level at<br />

10 m (peak SPL) Speed (mph) Motor size<br />

Idle 1 150 dB 5 115 hp<br />

2 151 dB 5 100 hp<br />

Planing 1 168 dB 35 115 hp<br />

2 163 dB 35 100 hp<br />

PWC 1 166 dB 40 1,235 cc<br />

2 158 dB 25 1,235 cc<br />

3 157 dB 25 1,235 cc<br />

4 162 dB 40 1,235 cc<br />

in conjunction with a Dell Inspiron 8100. The recording system had an overall<br />

frequency response of 20 Hz<strong>–</strong>22 kHz with a −<strong>17</strong>8 dB re 1 V/Pa sensitivity at 16bit<br />

resolution. For each stimulus, the sound was recorded as a vessel approached the<br />

recording hydrophone at a constant speed from approximately 500<strong>–</strong>1,000 m away.<br />

The vessel approached to exactly 10 m away from the recording hydrophone and<br />

continued away at a constant speed for another 500<strong>–</strong>1,000 m. This method enabled<br />

calculation of received levels of the watercraft at 10-m distance (Table 1). The 10-m<br />

distance was the closest point of approach. Acoustic recordings of vessel approaches<br />

were made at two boat speeds: idle and full-throttle planing. Multiple exemplars<br />

from each category were recorded in order to ensure the generality of results and<br />

reduce pseudoreplication, defined as generalization from a study due to an animal<br />

responding to or learning from a single exemplar (Kroodsma 1989). Characteristics<br />

of each exemplar are summarized in Table 1. All vessels recorded possessed a 4-stroke<br />

motor, which provided an element of standardization across playback categories,<br />

regardless of engine size or vessel type. Two vessels were recorded for idle approaches.<br />

One vessel had a 115-hp outboard, and the other vessel had 100 hp. The same vessels<br />

were used to record two exemplars during a planing approach. Two separate 4-stroke<br />

PWCs were used to produce two planing approach recordings from each vessel for a<br />

total of four exemplars. Each PWC was recorded at two planing speeds: 25 mph (40<br />

km/h) and 40 mph (64 km/h; Table 1).<br />

The duration of all stimulus exemplars, regardless of stimulus category, was edited<br />

to 3 min to ensure the same length for each exemplar. This was done to control<br />

for the amount of time the playback subject was exposed to any sound coming<br />

from the playback system. Three-minute durations were chosen based on the longest<br />

audible approach sequence, which was the slow moving idle approach. For the idle<br />

approaches, 45 s of ambient noise preceded and followed the onset and offset of the<br />

stimulus signal, resulting in a total exemplar duration of 3 min. For planing and<br />

PWC approaches, ambient noise preceding and following the stimulus signal was<br />

appropriately included to produce a final stimulus duration of 3 min. The duration<br />

of ambient recording added prior to the stimulus onset was equal to that added after<br />

the signal offset. A silent control of 3 min was also constructed by creating a null<br />

vector (sound file containing 3 min of silence) at the same sample rate as the other<br />

stimuli. A silent control was selected to expose the subjects to any extraneous noise<br />

added by the transmit system that could potentially elicit a reaction. A control of only<br />

background noise was not included in the playback sessions because ambient noise<br />

preceded boat noise stimuli in all stimulus categories. If the animals were significantly


MIKSIS-OLDS ET AL.: SIMULATED VESSEL APPROACHES 633<br />

Figure 1. Playback categories aligned in time. Top row (A) shows the acoustic envelope<br />

difference among the categories. The middle row (B) shows the raw time series of a single<br />

exemplar within each category. The bottom row (C) shows the corresponding spectrograms<br />

for each exemplar. Spectrograms are plotted on a relative dB scale. Received levels of each<br />

vessel type are presented in Table 1.<br />

reacting to the background noise, there would have been a significant reaction to all<br />

category approaches. There was no significant response to the idle approach, which<br />

indicated that the animals were not reacting to either the boat noise or the background<br />

noise.<br />

Playback Categories<br />

The exemplars in the playback categories varied in several acoustic parameters. The<br />

stimulus categories differ not only in their acoustic envelopes, or overall amplitude<br />

shape, but also in their frequency characteristics (Fig. 1). The idle approaches had<br />

the longest stimulus duration but the lowest stimulus amplitude (Fig. 1a, b). The<br />

planing approaches had a more rapid rise time and higher amplitude than the idle<br />

approaches, but with a shorter acoustic envelope. The planing approaches also had a<br />

more gradual onset compared to the abrupt offset. The acoustic envelope of the PWC<br />

approaches was the shortest with the most rise time and approximately equal peak<br />

amplitude as the planing approaches.<br />

The differences in frequency parameters were most evident in the spectrograms<br />

of Figure 1c. The idle approaches lacked a clearly defined broadband peak at the<br />

closest point of approach, and the U-shaped bands indicated the beta effect of sound<br />

during the approach and retreat (Tang 2005). The beta effect results in a decrease<br />

in signal frequency prior to the vessel’s closest point of approach and an increase<br />

in signal frequency during the retreat. From the perspective of the manatee, the<br />

beta effect could provide information about the position of the vessel. The planing<br />

and PWC approaches had a clear broadband peak at the closest point of approach.


634 MARINE MAMMAL SCIENCE, VOL. 23, NO. 3, 2007<br />

This was preceded by a strong tonal, harmonic signal. In the PWC approaches,<br />

the tonal component of the approach was vastly reduced compared to that of the<br />

planing approaches, and the broadband peak of the PWC approach was also much<br />

narrower. Figure 2 further illustrates the difference in frequency spectra among the<br />

playback categories. Fifteen seconds prior to the closest point of approach, the PWC<br />

was approximately 10 dB quieter than both the idle and planing signals at 2<strong>–</strong>3<br />

kHz (Fig. 2a). The magnitude of the idle and planing approaches was similar up<br />

to approximately 6 kHz. Above 6 kHz, the planing approach became about 10 dB<br />

louder than either the idle or PWC signals. Fifteen seconds prior to the closest<br />

point of approach, the planing approach transmitted the loudest signal in the higher<br />

frequencies. The planing approach was also clearly the loudest at the peak of approach<br />

by about 10 dB (Fig. 2b), with the exception of the PWC being the loudest between<br />

2 and 3 kHz. The idle approach was the quietest at all frequencies during the peak<br />

of approach.<br />

To better understand what the manatees detected during the vessel approaches,<br />

the power spectra of approaches were weighted by the only available manatee hearing<br />

thresholds as measured by Gerstein et al. (1999) (Fig. 3). Prior to the closest point of<br />

approach (Fig. 3a), planing signals were most salient above the background noise. The<br />

idle approach was above both the hearing threshold and noise floor for frequencies of<br />

approximately 2 kHz, whereas the PWC signature was either at or below detectable<br />

levels. At the closest point of approach (Fig. 3b), both the planing and PWC signals<br />

were well above threshold levels for all except the very lowest frequencies. The idle<br />

approach signal was loudest at approximately 2 kHz at the closest point of approach,<br />

but this level is 5 dB quieter than the planing and PWC levels over a band of 20 Hz<br />

to 20 kHz.<br />

Playback Experiments<br />

All playbacks were conducted in grass bed habitats that are commonly occupied<br />

by manatees. Restricting playbacks to one habitat type eliminated a confounding<br />

variable due to habitat type in the statistical analysis. In addition, playbacks were<br />

only performed with animals that were initially feeding or resting. The animals<br />

were relatively stationary during these two behaviors compared to traveling, milling,<br />

or social behaviors; therefore, changes in movement could be more easily observed<br />

and quantified. Restricting playbacks to two specific behavioral states also served to<br />

reduce the number of categories of analysis for statistical purposes, thus increasing<br />

the degrees of freedom within each category for each test.<br />

The playback protocol was designed to simulate a boat approaching a manatee to<br />

10 m. The same positional set-up was used for each playback regardless of which<br />

stimulus was used. The playback vessel was always positioned between the playback<br />

subject and the closest boating channel to simulate a boat approaching from deeper<br />

water and from the direction that a majority of boats would be traveling. Transmitted<br />

levels were adjusted for animal position so that the subject animal received<br />

appropriate received levels, as a function of stimulus type and exemplar (Table 1),<br />

regardless of the manatee’s distance to the playback vessel, which was between 2<br />

and 25 m from the manatee. Projection from a single, stationary transducer did not,<br />

however, allow for a directional motion component to the playback recording. If<br />

the manatees can localize the source accurately, the distance between the manatee<br />

and the playback vessel could have impacted the behavioral response. Due to the


A<br />

Power spectrum (dB re 1 Pa 2 /Hz)<br />

B<br />

Power spectrum (dB re 1 Pa 2 /Hz)<br />

130<br />

120<br />

110<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

MIKSIS-OLDS ET AL.: SIMULATED VESSEL APPROACHES 635<br />

Power Spectrum of Vessel Approach 15 s Before Peak<br />

30<br />

0 5 10 15 20 25<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

Frequency (Hz)<br />

Idle<br />

Plane<br />

PWC<br />

Power Spectrum for Vessel Approaches at Closest Point of Approach<br />

130<br />

Idle<br />

120<br />

Plane<br />

PWC<br />

110<br />

30<br />

0 5 10 15 20 25<br />

Frequency (Hz)<br />

Figure 2. Power spectra of vessel approach stimuli: (a) the comparison of 2-s clips measured<br />

15 s before the closest point of approach; (b) the comparison of 2-s clips taken at the closest<br />

point of approach.


636 MARINE MAMMAL SCIENCE, VOL. 23, NO. 3, 2007<br />

A<br />

Weighted Spectrum (dB)<br />

B<br />

Weighted Spectrum (dB)<br />

130<br />

120<br />

110<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

Weighted Spectrum for Vessel Approaches 15 s Before Peak<br />

30<br />

0 5 10 15 20 25 30 30<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

Frequency (kHz)<br />

Idle<br />

Plane<br />

PWC<br />

Noise<br />

Hearing Sensitivity<br />

Weighted Spectrum for Vessel Approaches at Closet Point of Approach<br />

130<br />

130<br />

120<br />

Idle<br />

Plane<br />

PWC<br />

120<br />

110<br />

Noise<br />

Hearing Sensitivity<br />

110<br />

100<br />

100<br />

30<br />

0 5 10 15 20 25 30 30<br />

Frequency (kHz)<br />

Figure 3. Spectra of vessel approaches weighted by the manatee hearing thresholds and<br />

smoothed in frequency plotted with ambient noise levels: (a) comparison of 2-s clips measured<br />

15 s before the closest point of approach; (b) comparison of 2-s clips taken at the closest point of<br />

approach. For illustration purposes the manatee hearing sensitivity values for the frequencies<br />

measured in this paper are shown by the dashed-dotted line (data taken from Gerstein et al.<br />

1999).<br />

130<br />

120<br />

110<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

Hearing Sensitivity (dB re 1 µPa)<br />

Hearing Sensitivity (dB re 1 µPa)


MIKSIS-OLDS ET AL.: SIMULATED VESSEL APPROACHES 637<br />

logistical constraints of the study design, this impact could not be addressed. Additional<br />

tests with stimuli simulating a rapid change in either direction or speed during<br />

the approach would provide valuable information that this study design was unable to<br />

offer.<br />

During playback experiments, the focal animal (occurring as a single animal or<br />

in a pair) was observed for a minimum of 20 min before and after the exposure of<br />

the playback stimuli. Primary observers were blind to the stimulus presentation, as<br />

the transducer and observer were on opposite sides of the boat and the background<br />

noise in the boat masked the in-air component of the playback stimulus, which was<br />

audible at the amplifier. During this time the subject was either identified from<br />

previously catalogued animals or photographed for later identification. Playbacks<br />

were only conducted if no other animals were detected within visual range during<br />

the 20-min pre-exposure period. Pre- and poststimulus observation included 4-min<br />

interval sampling of focal animal course, heading, distance to boat, and behavioral<br />

state. Ventilation and vocalization rates were recorded continuously. Vocalization<br />

rates were not analyzed, however, because the transducer was on the same vessel as<br />

the recording hydrophone, and the playback saturated the hydrophone recording rendering<br />

the manatee vocalizations inaudible. Saturated vocalization recordings during<br />

the playback sessions did not allow for uninterrupted vocalization rate analysis. If<br />

the animals left the area during the pre-exposure period or during exposure to the<br />

playback stimuli, they were not followed.<br />

Five minutes prior to the playback of any stimulus, the research vessel was anchored<br />

within 25 m of the focal animal and the sound source was deployed. An Aiwa model<br />

XP-V516C compact disc player connected to a Rockwood Detonator AMP-400 CRX<br />

amplifier delivered sound to a Lubell 9162 transducer, which projected the playback<br />

stimuli. This system was capable of producing a source level of approximately 190 dB<br />

re 1 Pa at 1 m in the frequency range of 240 Hz to 20 kHz. Response due to multiple<br />

boat interactions was avoided by only performing a playback when no other vessel<br />

had entered a 1-km radius for a 15-min period prior to the start of playback session.<br />

Each playback session consisted of four playback stimulus presentations presented<br />

5 min apart. The 5-min stimulus presentation was chosen based on the knowledge<br />

that bottlenose dolphins in Sarasota Bay encounter a passing vessel every 6 min<br />

(Buckstaff 2004). Manatees would experience a similar or shorter interval between<br />

boat encounters depending upon the number of recreational vessels in the area at any<br />

particular time. One stimulus from each of the four categories (control, idle, plane,<br />

PWC) was presented in random order. If the focal animal moved outside of a 25-m<br />

radius before the presentation of the last stimulus, the research vessel re-anchored<br />

closer to the animal, and the remaining playback stimuli were presented followed by<br />

a 20-min poststimulus observation period. In an effort to reduce pseudoreplication,<br />

no manatee was a playback subject more than twice, and no animal ever received the<br />

same exemplar more than once (Table 2).<br />

During the playback session, observations included point sampling of focal animal<br />

course, heading, distance to boat, ventilation, and behavior at the time of each<br />

surfacing. Visually observed responses to the playback stimuli generally fell into four<br />

categories: (1) investigate boat, (2) slow swim, (3) rolling dive, and (4) fast swim. Animals<br />

investigating the boat swam directly to the boat and interacted with it in some<br />

way. Slow swims were characterized by the animals changing position relative to the<br />

playback vessel without any visible wake or fluke prints. Rolling dives were identified<br />

by the arching of the manatee’s back and entire fluke leaving the water prior to a<br />

dive. Fast swims were characterized by a visible wake and strong fluke prints, often


638 MARINE MAMMAL SCIENCE, VOL. 23, NO. 3, 2007<br />

Table 2. Summary of playback subjects, stimulus sequence, and response. In the “age” and “group composition” categories, A indicates adult, SA<br />

indicates subadult, C indicates calf<br />

Number Group Stimulus 1 Stimulus 2 Stimulus 3 Stimulus 4<br />

Session Date ID Age in Group composition Response Response Response Response<br />

1 4 June 2004 Snuff A 2 A, A C Idle2 PWC4 —<br />

(SB047) None None FS, LA —<br />

2 10 June 2004 Phish LA A 1 Plane2 C Idle2 PWC2<br />

SS None SS RD<br />

3 15 June 2004 Phish A 1 PWC2 Plane1 Idle2 C<br />

SS None None None<br />

4 21 June 2004 DU723 A 2 A, C Plane2 Idle2 C PWC1<br />

SS None None FS<br />

5 23 June 2004 623BB A 2 A, A C Idle1 Plane1 PWC3<br />

None SS FS FS, LA<br />

6 25 June 2004 625A A 2 A, C PWC4 Idle2 Plane2 C<br />

FS SS SS None<br />

7 28 June 2004 628A A 2 A, C PWC1 C Plane1 Idle1<br />

FS SS RD None<br />

8 1 July 2004 DU723 A 2 A, C C Idle1 PWC2 Plane1<br />

None SS FS SS<br />

9 6 July 2004 706AA SA 1 PWC4 Idle1 C Plane1<br />

RD None None None<br />

10 21 July 2004 721A A 2 A, C Idle1 Plane2 C PWC1<br />

None None None FS, LA


Table 2. Continued.<br />

MIKSIS-OLDS ET AL.: SIMULATED VESSEL APPROACHES 639<br />

11 22 July 2004 Clyde A 2 A, A C PWC4 Plane2 Idle2<br />

None SS None None<br />

12 23 July 2004 723B A 2 A, C Plane2 C PWC3 Idle1<br />

RD None RD RD<br />

13 23 July 2004 723A A 2 A, C PWC2 C Plane2 —<br />

RD None SS, LA —<br />

14 26 July 2004 726D A 2 A, C Plane1 Idle2 PWC1 C<br />

FS SS FS None<br />

15 26 July 2004 Phish LA A 2 A, A C Idle1 PWC4 Plane1<br />

None None FS SS, LA<br />

16 27 July 2004 Splotch A 2 A, A PWC2 Idle1 C Plane1<br />

FS None None None<br />

<strong>17</strong> 28 July 2004 727AA SA 1 C Idle2 — —<br />

SS None — —<br />

18 3 August 2004 803BB A 2 A, SA Plane2 C Idle1 PWC3<br />

RD None RD SS, LA<br />

19 4 August 2004 804C SA 2 SA, SA C PWC4 Idle2 Plane1<br />

None FS None None<br />

20 5 August 2004 805BB A 2 A, A Idle1 Plane1 PWC1 C<br />

None FS FS SS<br />

21 19 August 2004 823F SA 1 Idle2 PWC2 Plane1 C<br />

None SS None SS


640 MARINE MAMMAL SCIENCE, VOL. 23, NO. 3, 2007<br />

accompanied by a mud cloud stirred up from the bottom. Two retreating behaviors<br />

were also observed in relation to the playback vessel: (1) retreat to deep water and<br />

(2) pass by boat on the way to deep water. Direct retreats were characterized by the<br />

animal increasing distance from the playback vessel in the direction of deeper water.<br />

Passing by the boat on the way to deep water was a separate classification because<br />

the animal had to initially approach the playback vessel before retreating to deeper<br />

water. Passing by the boat on the way to deep water differed from the “investigate<br />

boat” response because the passing by the boat did not include any direct contact or<br />

interaction with the playback vessel by the manatee.<br />

In addition to behavioral responses, a respiratory response to the playback stimuli<br />

was also calculated. An index of ventilation variability (index of variability) was<br />

calculated from the time series of visually observed surface breaths recorded continuously<br />

throughout the playback experiments. The poststimulus index of variability<br />

was calculated by adding the absolute values of the difference in time between breaths<br />

for the three consecutive breaths following the stimulus onset (poststimulus index<br />

of variability =|t 0 − t 1| +|t 1 − t 2| +|t 2 − t 3|). Higher values of poststimulus<br />

variability indicated a larger variability in ventilation. Traditional measures of ventilation<br />

rate and its associated variance were not utilized in this analysis because long<br />

submersions followed by quick successive breaths cancelled each other and masked<br />

the true breathing pattern. Inclusion of absolute values preserved the level of variability<br />

between breaths. Prestimulus index of variability was calculated in a similar<br />

manner. The index was calculated from a time series of triads consisting of three consecutive<br />

breaths preceding the first stimulus. The triad values were then averaged to<br />

produce a prestimulus index of variability (prestimulus index of variability ={(|p 0<br />

− p 1|+|p 1 − p 2|+|p 2 − p 3|)}/n). Multiple, consecutive triads were used in the<br />

prestimulus index in order to establish the most accurate baseline index of variability.<br />

The poststimulus index only contained a single triad because this was the minimum<br />

number of surfacings observed between stimulus presentations during the playback<br />

session.<br />

Statistical Analysis<br />

The first level of analysis compared the frequency of animals either responding<br />

or not responding to each playback category and the control condition. Statistical<br />

significance was based on the binomial distribution. Differences in the frequency of<br />

behavioral responses between categories were investigated with a R × C G-test for<br />

independence. Based on behavioral observations to opportunistic vessel approaches,<br />

the three categories of response (slow swim, fast swim, and rolling dive) were defined<br />

prior to the initiation of the playback study. A priori conditions were met, which<br />

justified the use of the R × C G-test for independence used to assess the frequency<br />

of behavioral responses to the playback stimuli. Differences in the deviation between<br />

prestimulus and poststimulus index of variability across playback categories were<br />

tested with a single-factor analysis of variance (ANOVA).<br />

RESULTS<br />

Twenty-one playback sessions were conducted with 19 different subjects encountered<br />

either as single animals or in pairs (Table 2). A total of eighty stimuli were<br />

played: 21 Control, 10 Idle1, 10 Idle2, 11 Planing1, 8 Planing2, 6 PWC1, 5 PWC2,<br />

3 PWC3, and 6 PWC4. Seventeen out of 21 animals (81%) showed no locomotor


MIKSIS-OLDS ET AL.: SIMULATED VESSEL APPROACHES 641<br />

Figure 4. Proportion of responses elicited by playback treatments.<br />

response to the silent control (Table 2 and Figure 4). The difference between response<br />

or no response for all stimuli vs. the silent control was significant at the 95% significance<br />

level based on a binomial distribution (P < 0.001). This indicated that the<br />

manatees were not significantly reacting to the exposure of the broadcasting system<br />

provided by the control stimulus. Manatees showed a marked locomotor response<br />

to the playback stimuli compared to the silent control. Thirteen of twenty animals<br />

(65%) showed no response to the idle approach, whereas 35% showed some type<br />

of locomotor response (i.e., slow swim, approach boat, fast swim, etc.) (Figure 4).<br />

This 35% response rate was not significant at the 95% significance level. During<br />

the planing approaches there was a significant locomotor response rate of 63% (P =<br />

0.002). Of the twelve animals that showed a locomotor response to the planing approach,<br />

two abandoned the area. All animals showed a locomotor response to the<br />

PWC approach (P < 0.001). Four of the twenty animals that responded (20%) left<br />

the area.<br />

An analysis of response orientation and heading of those animals that did show<br />

a locomotor response to the playback stimuli revealed a striking pattern (Fig. 4).<br />

Of the four animals that did respond to the control, all four investigated the boat.<br />

Seven animals of twenty responded to the idle approach. Of the seven, four (57%)<br />

retreated directly to deep water, two (29%) passed by the playback vessel on the way<br />

to deeper water, and one animal (14%) retreated from the playback vessel to shallow<br />

water. During the planing and PWC approaches, the number of animals retreating<br />

directly to deep water increased, whereas the number of animals passing by the boat<br />

decreased. In general, manatees tended to respond to all approaches by retreating to<br />

deep water. The frequency of animals retreating directly to deep water increased in<br />

response to an increase in speed of the approaching vessel.<br />

Behavioral analysis of the retreating animals showed a graded response in behavior<br />

associated with playback category (Fig. 5). No animals retreated during the controls,<br />

so this category was not included in the analysis. The frequency of animals retreating<br />

with a slow swim decreased from 71% in response to the idle approach to 37% for the


642 MARINE MAMMAL SCIENCE, VOL. 23, NO. 3, 2007<br />

Figure 5. Behavioral response of retreating manatees to the playback categories. There<br />

is no response for the controls because no animals retreated during control presentations. ∗<br />

shows the significant decrease in the slow swim response associated with playback category<br />

(0.01 < P < 0.025). ∗∗ shows the significance increase in fast swim responses associated with<br />

playback category (0.01 < P < 0.025).<br />

planing approach to finally 16% in response to the PWC approach (Fig. 5a). Analysis<br />

of frequency using a R × C G-test for independence revealed that the frequency of<br />

the slow swim response was dependent on playback category (0.01 < P < 0.025;<br />

Sokal and Rohlf 1995). Similarly, the increase seen in the frequency of fast swim<br />

response was dependent upon playback category (0.01 < P < 0.025). No animals<br />

responded to the idle approach with a fast swim, whereas 37% and 68% responded<br />

to the planing and PWC approaches, respectively (Fig. 5b). There was slight decrease<br />

in the frequency of rolling dive responses to the PWC approaches, but this decrease<br />

was not a significant pattern.<br />

An ANOVA of the poststimulus index of variability showed a significant overall<br />

effect of stimulus type (F3,82 = 2.72, P = 0.04; Fig. 6). Post hoc multiple comparisons<br />

indicated an increase in variability between the prestimulus and both planing<br />

Figure 6. Mean deviation from prestimulus/control variability for each playback category.<br />

The asterisk ( ∗ ) indicates categories that differed significantly from the prestimulus/control<br />

values. Error bars represent standard error.


MIKSIS-OLDS ET AL.: SIMULATED VESSEL APPROACHES 643<br />

and PWC responses. Post hoc multiple comparisons were performed for all other<br />

category comparisons, but no other comparisons were significant. This suggests that<br />

the manatees responded to the planing and PWC approaches in a similar manner<br />

with an increase in ventilation variability.<br />

DISCUSSION<br />

The most pronounced responses to the playback stimuli, relative to the controls,<br />

were elicited by the PWC. Significant behavioral and physiological responses were also<br />

seen in response to planing boat approaches, indicating that rapid vessel approaches do<br />

affect manatee behavior. Approaches by fast-moving vessels resulted in the disruption<br />

of feeding activity, an increase in energy expenditure inferred from swim speed, and<br />

in some cases a short-term avoidance of the feeding area. Avoidance reactions to<br />

approaching vessels are not unique to manatees, as disturbance responses to motorized<br />

vehicles have been documented in both marine and terrestrial species. In the marine<br />

environment, avoidance to motorized watercraft had been reported in manatees,<br />

cetaceans, and pinnipeds (manatees: Provancha and Provancha [1988], Buckingham<br />

et al. [1999], Nowacek et al. [2004a]; cetaceans: bottlenose dolphins, Janik and<br />

Thompson [1996], Nowacek et al. [2001b], Hastie et al. [2003], Buckstaff [2004];<br />

killer whales, Orcinus orca, Kruse [1991], Williams et al. [2002a, b]; Hector’s dolphins,<br />

Cephalorhynchus hectori, Bejder et al. [1999]; beluga whales, Delphinapterus leucas, Finley<br />

et al. [1990]; pinnipeds: walruses, Odobenus rosmarus, Fay et al. [1984]; and harbor<br />

seals, Phoca vitulina, Reijnders [1981]). Documented disturbance reactions include<br />

increases in vocalization rate, increases in swim speed, longer dive durations, decreased<br />

interanimal distance, increased breathing synchrony, and displacement from haulout<br />

sites. Terrestrial animals (bighorn sheep, Ovis canadensis canadensis, MacArthur<br />

et al. [1979]; white-tailed deer, Odocoileus virginianus, Richens and Lavigne [1978];<br />

caribou, Rangifer tarandus, Murphy et al. [1993]; and penguins, Pygoscelis adelia, Culik<br />

et al. [1990]) were also found to avoid road vehicles, snowmobiles, and aircraft.<br />

The manatees studied here showed the ability to discriminate and differentially<br />

react to the two different engine types and speeds simulated in the playback experiments.<br />

Nowacek et al. (2001a, 2004a) reported a generalized response by manatees<br />

to approaching boats involving turning toward or into deep water without specific<br />

regard to boat type, boat speed, distance from the manatee, the kind of habitat the<br />

boat was operating in, or the kind of habitat occupied by the manatee. Increases in<br />

swim speed were most prevalent in shallow water grassbeds when boats approached<br />

to between 0 and9m(Nowacek et al. 2004a). This study also showed that manatees<br />

reacted to simulated vessel approaches to within 10 m with an increase in swim<br />

speed and directed movement toward the closest deep water. The findings here differed<br />

from the previous study, however, because boat type and boat speed in this<br />

study appear to have a significant effect on swimming speed. This effect may not<br />

have been detected by Nowacek et al. (2004a) due to differences in study design and<br />

categorization of visible responses. This study differentiated between responses based<br />

on two different changes in swim speeds and the presence of rolling dives, which<br />

indicate deeper dives. Results presented here show that the manatees responded to<br />

slower idle approaches with a greater number of slow swim responses and a larger<br />

number of retreat paths that intersected with the playback vessel. In contrast, responses<br />

to fast approaching outboard motorboats or PWCs elicited a significantly<br />

greater frequency of response for fast swim speeds and retreat paths that avoided


644 MARINE MAMMAL SCIENCE, VOL. 23, NO. 3, 2007<br />

the playback vessel. Because playback experiments only introduce the frequency and<br />

amplitude components of vessel signals (other potential components include visual<br />

information above and below water, and changes in the bearing of the acoustic signal),<br />

the frequency and amplitude information available to the animal prior to the closest<br />

point of approach can be used to explain how the animals may be discriminating and<br />

ultimately reacting to the different playback categories.<br />

Fifteen seconds prior to the closest point of approach, the planing approaches were<br />

approximately 10 dB louder than the idle and PWC approaches for frequencies from<br />

6to22kHz. Similarly, the idle approach was approximately 12 dB louder than both<br />

the planing and PWC approaches at 2 kHz. Therefore, the slower rise times of the<br />

idle and planing approaches provide more information to the animals 15 s prior to<br />

arrival compared to the PWC approach. It is possible the manatees can extract the<br />

necessary information from these acoustic cues relating to speed, direction, and boat<br />

type in order to execute the most appropriate and energetically favorable response.<br />

The most desirable responses are those that enable manatees to avoid collisions with<br />

watercraft. The playbacks provided an opportunity to observe responses to specific<br />

acoustic components of vessels approaching at a constant speed under controlled<br />

circumstances. Responses indicated that manatees are able to detect approaching<br />

vessels and execute appropriate responses to avoid vessel collisions by retreating to<br />

deep water where they can swim beneath the boat. Furthermore, manatees are able to<br />

discriminate between vessel types and speeds, which appeared to influence the degree<br />

and direction of response. Manatees tended not to respond to idle approaches. When<br />

animals did respond, most did so at a slow speed. By contrast, responses to PWC and<br />

planning approaches elicited a greater proportion of fast swims. Slower responses to<br />

slowly approaching vessels and quicker responses to faster-moving vessels would be<br />

appropriate in real-life situations, provided the vessels did not change speed. Manatee<br />

response is in marked contrast to that of right whales, which tend not to respond<br />

to approaching ships even though they do respond to alert signals (Nowacek et al.<br />

2004b). Right whales responded to an alert signal during controlled exposures by<br />

swimming strongly to the surface, which is a response that is likely to increase rather<br />

than decrease the risk of collision.<br />

The most energetically favorable response for a manatee to any vessel approach<br />

would be to minimize locomotor costs by not moving at all. However, this would<br />

not always be the most appropriate response in terms of avoiding vessel collisions. If<br />

a change in location is necessary, a response at swim speeds at or near the minimum<br />

cost of transport would be most efficient. It is often found that birds and mammals<br />

swim underwater at or near the speed of minimum cost of transportation (Williams et<br />

al. 1993, Ropert-Coudert et al. 2001, Lavvorn et al. 2004). Manatees generally cruise<br />

at speeds of 2<strong>–</strong>6 mph (3<strong>–</strong>10 km/h), although they have been recorded at speeds of<br />

15 mph (24 km/h) for short bursts (Hartman 1979). Speeds of 2<strong>–</strong>6 mph would have<br />

been classified into the slow swim response in this study, so it appears that manatees<br />

responding to idle and many planing approaches acquire enough prior information to<br />

execute an energetically efficient response. The PWC acoustic signatures 15 s prior<br />

to arrival do not provide as much acoustic information compared to the idle and<br />

planing approaches.<br />

The short rise time signal associated with PWC approaches does not differ greatly<br />

from ambient noise levels until 5 s before the peak, so it is possible that the<br />

manatees do not perceive these approaches in enough time to execute an energetically<br />

favorable response. Consequently, faster, less efficient responses are necessary to


MIKSIS-OLDS ET AL.: SIMULATED VESSEL APPROACHES 645<br />

retreat from a possible PWC collision. An alternative explanation is that the sharp<br />

rise time associated with the PWC approach elicits a startle response that causes<br />

manatees to retreat from the sound source without evoking a higher level of cognitive<br />

analysis. Avoidance responses to the short rise time signals have also been<br />

observed in sharks. Myrberg et al. (1978) reported that a silky shark (Carcharhinus<br />

falciformis) withdrew 10 m from a speaker broadcasting a 150<strong>–</strong>600-Hz sound<br />

with a sudden onset and a peak sound pressure level of 154 dB re 1 Pa. These<br />

sharks also avoided a pulsed attractive sound when its sound level was abruptly increased<br />

by >20 dB. Finally, through prior encounters with PWCs, manatees may<br />

have learned to associate the PWC acoustic signal with vessels that are less predictable<br />

and more likely to approach them in shallow areas, thus requiring a more extreme<br />

response.<br />

Regardless of the specific acoustic characteristic of the fast vessel approaches that<br />

elicit fast swim responses, these signals cause manatees to increase their swim speed.<br />

Swim speed, as well as breathing rates and heart rate, have been used to estimate the<br />

energies of free-ranging marine mammals (Sumich 1983, Kshatriya and Blake 1988,<br />

Williams et al. 1992, Hind and Gurney 1997). Assuming that manatees respond at<br />

maximum speeds and that the maximum aerobic energy used during locomotion can<br />

reach 4<strong>–</strong>11 times resting levels in marine mammals (Elsner 1986, Williams et al.<br />

1993), consistent responses to vessel approaches potentially affect the energy budget of<br />

manatees in a significant way. Responses detected in this study are consistent with an<br />

energy cost, and future work quantifying energy expenditures will determine whether<br />

multiple reactions could have a long-term effect at the individual or population<br />

level.<br />

Most playback experiments measure movement, visual, or vocal responses to the<br />

sound played. Behaviors most typically measured are orientation or movement relative<br />

to the sound source, vocalizations made in response to the playback, and previously<br />

defined behaviors or displays such as aggressive or sexual displays. Less frequently<br />

used, but possibly more objective, responses are changes in heart rate (birds: Davis<br />

[1986], Diehl [1992]; humans: Brown et al. [1976]; chimpanzees, Pan troglodytes:<br />

Berntson and Boysen [1989]; bottlenose dolphins: Miksis et al. [2001]) and hormone<br />

levels (Dufty 1982). Neither quantitative swim speed nor fluke rate or amplitude<br />

was measurable in this study. More accurate measurements of both swimming characteristics<br />

and physiological responses during playback responses are necessary in<br />

order to determine the degree to which repeated exposure to vessel approaches are<br />

affecting the manatee energy budget or stress levels. Technological advances in tag<br />

construction and measurement sensors may soon allow for the recording of these<br />

critical parameters (Johnson and Tyack 2003).<br />

In summary, the playback technique presented here permits the investigation of<br />

numerous questions associated with manatee disturbance, threshold level, etc. without<br />

the risk of injury associated with the unpredictable behavior of wild animals<br />

during directed vessel approaches. This methodology has identified that vessel approaches,<br />

especially by PWCs and fast approaching watercraft, are a cause of manatee<br />

disturbance. This may be of regulatory concern, as harassment is prohibited in the<br />

Unites States by the Marine Mammal Protection Act. Manatees were also shown to<br />

hear and respond to boats approaching at idle speeds. Much more information is<br />

needed to determine how to minimize this disturbance in order to meet the criteria<br />

for species downlisting as outlined in the Florida Manatee Recovery Plan (U.S. Fish<br />

and Wildlife Service 2003). For example, the measured responses in this study were


646 MARINE MAMMAL SCIENCE, VOL. 23, NO. 3, 2007<br />

only made during feeding and resting behaviors in grass bed habitats. Behavioral<br />

state and habitat type may have significant impacts on the motivation to react, and<br />

responses in other behavioral contexts (e.g., actively socializing) and habitats should<br />

be investigated.<br />

ACKNOWLEDGMENTS<br />

The success of this project would not have been possible without the help of numerous<br />

people. The interns and staff in the Manatee Research Program at Mote Marine Laboratory were<br />

critical to the field efforts. Special thanks are extended to Doug Nowacek for his input on initial<br />

project design. Thanks are also extended to Gretchen Hurst, Joe Gaspard, and Marie Chapla<br />

for their help with stimulus recordings. All playbacks were conducted under the guidelines<br />

of U.S. Fish and Wildlife permit MA07<strong>17</strong>99-0 issued to Jennifer Miksis and comply with all<br />

current laws of the country. This project was funded by a National Defense for Science and<br />

Engineering Graduate (NDSEG) Fellowship, P.E.O. Scholar Award, and American Association<br />

for University Women (AAUW) Dissertation Writing Fellowship awarded to Jennifer Miksis.<br />

LITERATURE CITED<br />

ACKERMAN, B.B.,S. D. WRIGHT, R.K.BONDE, D.K.ODELL and D. J. BANOWETZ.<br />

1995. Trends and patterns in mortality of manatees in Florida. Pages 1974<strong>–</strong>1992 in T.<br />

J. O’Shea, B. B. Ackerman and H. F. Percival, eds. Population biology of the Florida<br />

manatee (Trichechus manatus latirostris). National Biological Service, Information and<br />

Technical Report 1. Washington, DC.<br />

BEJDER, L., S. DAWSON AND J. A. HARRAWAY. 1999. Responses by Hector’s dolphins to<br />

boats and swimmers in Porpoise Bay, New Zealand. Marine Mammal Science 15:738<strong>–</strong><br />

750.<br />

BENGTSON,J.L.,AND S. M. FITZGERALD. 1985. Potential role of vocalizations in West Indian<br />

manatees. Journal of Mammalogy 66:816<strong>–</strong>819.<br />

BERNTSON,G.G.,AND S. T. BOYSEN. 1989. Specificity of the cardiac response to conspecific<br />

vocalizations in chimpanzees. Behavioral Neuroscience 103:235<strong>–</strong>243.<br />

BROWN, J.W.,P.A.MORSE, L.A.LEAVITT AND F. K. GRAHAM. 1976. Specific attentional<br />

effects reflected in the cardiac orienting response. Bulletin of the Psychonomic Society<br />

7:1<strong>–</strong>4.<br />

BUCKINGHAM,C.A.,L. W. LEFEBVRE,J.M.SCHAEFER AND H. I. KOCHMAN. 1999. Manatee<br />

response to boating activity in a thermal refuge. Wildlife Society Bulletin 27:514<strong>–</strong>522.<br />

BUCKSTAFF, K.C.2004. Effects of watercraft noise on the acoustic behavior of bottlenose<br />

dolphins, Tursiops truncatus, inSarasota Bay, Florida. Marine Mammal Science 20:709<strong>–</strong><br />

725.<br />

CULIK,B.M.,D. ADELUNG AND A. J. WOAKES. 1990. Effects of disturbance on the heart rate<br />

and behavior of Adelie penguins (Pygoscelis adeliae) during the breeding season. Pages<br />

<strong>17</strong>7<strong>–</strong>282 in K. R. Kerry and G. HEMPEL, eds. Antarctic ecosystems. Ecological change<br />

and conservation. Springer, New York, NY.<br />

DAVIS,W.J.1986. Acoustic recognition in the belted kingfisher: Cardiac responses to playback<br />

vocalizations. Condor 88:505<strong>–</strong>512.<br />

DIEHL, P.1992. Radiotelemetric measurements of heart rate in singing blackbirds (Turdus<br />

merula). Ornithology 133:181<strong>–</strong>195.<br />

DUFTY,A.M.1982. Responses of brown-headed cowbirds to simulated conspecific intruders.<br />

Animal Behaviour 30:1043<strong>–</strong>1052.<br />

ELSNER,R.1986. Limits to exercise performance: Some ideas from comparative studies. Acta<br />

Physiologica Scandinavica 128:45<strong>–</strong>51.<br />

FAY, F.H.,B. P. KELLY, P.H.GEHNRICH, J.L.SEASE AND A. A. HOOVER. 1984. Modern<br />

populations, migrations, demography, trophics, and historical status of the Pacific walrus.


MIKSIS-OLDS ET AL.: SIMULATED VESSEL APPROACHES 647<br />

Pages 231<strong>–</strong>376 in Outer Continental Shelf Environmental Assessment Program, Final<br />

Report 693. NOAA, Anchorage, AK.<br />

FINLEY, K.J.,G. W. MILLER, R.A.DAVIS AND C. R. GREENE. 1990. Reactions of belugas<br />

and narwhals to ice breaking ships in the Canadian high arctic. Canadian Bulletin of<br />

Fisheries Aquatic and Sciences 224:97<strong>–</strong>1<strong>17</strong>.<br />

GERSTEIN,E.R.2002. Manatees, bioacoustics and boats. American Scientist 90:154<strong>–</strong>163.<br />

GERSTEIN, E.R.,L. GERSTEIN, S.E.FORSYTHE AND J. E. BLUE. 1999. The underwater<br />

audiogram of the West Indian manatee (Trichechus manatus). Journal of the Acoustical<br />

Society of America 105:3575<strong>–</strong>3583.<br />

GORZELANY, J.F.2004. Evaluation of boater compliance with manatee speed zones along<br />

the Gulf coast of Florida. Coastal Management 32:215<strong>–</strong>226.<br />

HARTMAN, D.S.1979. Ecology and behavior of the manatee (Trichechus manatus) in Florida.<br />

Special Publication 5, American Society of Mammalogists, Lawrence, KS.<br />

HASTIE, G.D., B. WILSON, L.H.TUFFT AND P. M. THOMPSON. 2003. Bottlenose dolphins<br />

increase breathing synchrony in response to boat traffic. Marine Mammal Science 19:74<strong>–</strong><br />

84.<br />

HIND, A.T.,AND W. S. C. GURNEY. 1997. The metabolic cost of swimming in marine<br />

homeotherms. Journal of Experimental Biology 200:531<strong>–</strong>542.<br />

JANIK, V.M.,AND P. M. THOMPSON. 1996. Changes in surfacing patterns of bottlenose<br />

dolphins in response to boat traffic. Marine Mammal Science 12:597<strong>–</strong>602.<br />

JOHNSON, M.P.,AND P. L. TYACK. 2003. A digital acoustic recording tag for measuring<br />

the response of wild marine mammals to sound. IEEE Journal of Ocean Engineering<br />

28:3<strong>–</strong>12.<br />

KETTEN,D.R.,D. K. ODELL AND D. P. DOMNING. 1992. Structure, function, and adaptation<br />

of the manatee ear. Pages 77<strong>–</strong>95 in J. A. Thomas, R. A. Kaestelein, and A. Y. Supin,<br />

eds. Marine mammal sensory systems. Plenum Press, New York, NY.<br />

KROODSMA, D.E.1989. Suggested experimental designs for song playbacks. Animal Behaviour<br />

37:600<strong>–</strong>609.<br />

KRUSE, S.1991. The interactions between killer whales and boats in Johnstone Strait, B.C.<br />

Pages 149<strong>–</strong>159 in K. Pryor and K. S. NORRIS, eds. Dolphin societies: Discoveries and<br />

puzzles. University of California Press, Berkeley, CA.<br />

KSHATRIYA, M.,AND R. W. BLAKE. 1988. Theoretical model of migration energetics in the<br />

blue whale, Balaenoptera musculus. Journal of Theoretical Biology 133:479<strong>–</strong>498.<br />

LAVVORN,J.R.,Y. WATANUKI,A.KATO,Y.NAITO AND G. A. LIGGINS. 2004. Stroke patterns<br />

and regulation of swim speed and energy cost in free-ranging Brunnich’s guillemots.<br />

Journal of Experimental Biology 207:4679<strong>–</strong>4695.<br />

MACARTHUR, R.A.,R. H. JOHNSON AND V. GEIST. 1979. Factors influencing heart rate in<br />

free-ranging bighorn sheep: A physiological approach to the study of wildlife harassment.<br />

Canadian Journal of Zoology 57:2010<strong>–</strong>2021.<br />

MEDWIN, H., AND C. S. CLAY. 1998. Fundamentals of acoustical oceanography. Academic<br />

Press, New York, NY.<br />

MIKSIS, J.L.,M. D. GRUND, D.P.NOWACEK, A.R.SOLOW, R.C.CONNOR AND P.<br />

L. TYACK. 2001. Cardiac responses to acoustic playback experiments in the captive<br />

bottlenose dolphin (Tursiops truncatus). Journal of Comparative Psychology 115:227<strong>–</strong><br />

232.<br />

MURPHY, S.M.,R. G. WHITE, B.A.KUGLER, J.A.KITCHENS, M.D.SMITH AND D. S.<br />

BARBER. 1993. Behavioral effects of jet aircraft on caribou in Alaska. Proceedings of the<br />

6th <strong>International</strong> Congress Noise as a Public Health Problem 3:479<strong>–</strong>486.<br />

MYRBERG, Jr.,A.A.,C.R.GORDON AND A. P. KLIMLEY. 1978. Rapid withdrawal from<br />

a sound source by open-ocean sharks. Journal of the Acoustical Society of America<br />

64:1289<strong>–</strong>1297.<br />

NOWACEK,S.M.,R. S. WELLS,D.P.NOWACEK,E.C.G.OWEN,T.R.SPEAKMAN AND R. O.<br />

FLAMM. 2000. Manatee behavioral responses to vessel approaches. Final contract report<br />

to the Florida Fish and Wildlife Conservation Commission. Mote Marine Laboratory<br />

Technical Report No. 742. Sarasota, FL. 12 pp. + appendices.


648 MARINE MAMMAL SCIENCE, VOL. 23, NO. 3, 2007<br />

NOWACEK,S.M.,R. S. WELLS,D.P.NOWACEK,E.C.G.OWEN,T.R.SPEAKMAN AND R. O.<br />

FLAMM. 2001a. Manatee behavioral response to vessel traffic. 14th Biennial Conference<br />

on the Biology of Marine Mammals. Vancouver, BC, Canada, 28 Nov, 3 Dec, 2001.<br />

NOWACEK,S.M.,R. S. WELLS AND A. R. SOLOW. 2001b. Short-term effects of boat traffic on<br />

bottlenose dolphins, Tursiops truncatus,inSarasota Bay, Florida. Marine Mammal Science<br />

<strong>17</strong>:673-688.<br />

NOWACEK, D.P.,B.M.CASPER, R.S.WELLS, S.M.NOWACEK AND D. A. MANN. 2003.<br />

Intraspecific and geographic variation of West Indian manatee (Trichechus manatus spp.)<br />

vocalizations (L). Journal of the Acoustical Society of America 114:66<strong>–</strong>69.<br />

NOWACEK, S.M.,R. S. WELLS, E.C.G.OWEN, T.R.SPEAKMAN, R.O.FLAMM AND<br />

D. P. NOWACEK. 2004a. Florida manatees, Trichechus manatus latirostris, respond to<br />

approaching vessels. Biological Conservation 119:5<strong>17</strong><strong>–</strong>523.<br />

NOWACEK, D.P.,M.P.JOHNSON AND P. L. TYACK. 2004b. North Atlantic right whales<br />

(Eubalaena glacialis) ignore ships but respond to alerting stimuli. Proceedings of the<br />

Royal Society of London B 271:227<strong>–</strong>231.<br />

PROVANCHA, J.A.,AND M. J. PROVANCHA. 1988. Long-term trends in abundance and<br />

distribution of manatees in the northern Banana River, Brevard County, FL. Marine<br />

Mammal Science 4:323<strong>–</strong>338.<br />

REIJNDERS, P.J.H.1981. Management and conservation of the harbour seal, Phoca vitulina,<br />

population in the international Wadden Sea area. Biological Conservation 19:213<strong>–</strong>221.<br />

REYNOLDS,J.E., III. 1981. Behavior patterns in the West Indian manatee, with emphasis on<br />

feeding and diving. Florida Scientist 44:233<strong>–</strong>242.<br />

REYNOLDS, J.E.1999. Efforts to conserve manatees. Pages 267<strong>–</strong>295 in J.R. Twiss and R.R.<br />

Reeves, eds. Conservation and management of marine mammals. Smithsonian Institution<br />

Press, Washington, DC.<br />

RICHARDSON, W.,C. GREENE, C.MALME AND D. THOMSON. 1995. Marine mammals and<br />

noise. Academic Press, San Diego, CA.<br />

RICHENS, V.B.,AND G. R. LAVIGNE. 1978. Response of white-tailed deer to snowmobiles<br />

and snowmobile trails in Maine. Canadian Field-Naturalist 92:334<strong>–</strong>344.<br />

ROPERT-COUDERT,Y.,K. SATO,J.BAUDAT,C.A.BOST,Y.LE MAHO AND Y. NAITO. 2001.<br />

Feeding strategies of free-ranging Adelie penguins Pygoscelis adeliae analysed by multiple<br />

data recording. Polar Biology 24:460<strong>–</strong>466.<br />

SOKAL,R.R.,AND F. J. ROHLF. 1995. Biometry. 3rd edition. W. H. Freeman and Company,<br />

New York, NY.<br />

SUMICH, J.L.1983. Swimming velocities, breathing patterns, and estimated costs of locomotion<br />

in migrating gray whales, Eshrichitus robustus. Canadian Journal of Zoology<br />

61:647<strong>–</strong>652.<br />

TANG, D.2005. Inverting for sandy sediment sound speed in very shallow water using boat<br />

noise. Journal of the Acoustical Society of America 118:2503.<br />

U.S. FISH AND WILDLIFE SERVICE. 2003. Florida manatee recovery plan, (Trichechus manatus<br />

latirostris), 3rd revision. U.S. Fish and Wildlife Service, Atlanta, GA.<br />

URICK, R.J.1983. Principles of underwater sound. 3rd edition. Peninsula Publishing, Los<br />

Altos, CA.<br />

WEIGLE,B.L., I. E. WRIGHT AND J. A. HUFF. 1994. Responses of manatees to an approaching<br />

boat: A pilot study. First <strong>International</strong> Manatee and Dugong Research Conference, 11<strong>–</strong><br />

13 March 1994, Gainesville, FL.<br />

WELLS, R.S.,AND M. D. SCOTT. 1997. Seasonal incidence of boat strikes on bottlenose<br />

dolphins near Sarasota, Florida. Marine Mammal Science 13:475<strong>–</strong>480.<br />

WILLIAMS, T.M.,W. A. FRIEDL, M.L.FONG, R.M.YAMADA, P.SEDIVY AND J. E. HAUN.<br />

1992. Travel at low energetic cost by swimming and wave-riding of bottlenose dolphins.<br />

Nature 355:821<strong>–</strong>823.<br />

WILLIAMS, T.M.,W.A.FRIEDL AND J. E. HAUN. 1993. The physiology of the bottlenose<br />

dolphin (Tursiops truncatus): Heart rate, metabolic rate, and plasma lactate concentration<br />

during exercise. Journal of Experimental Biology <strong>17</strong>9:31<strong>–</strong>46.


MIKSIS-OLDS ET AL.: SIMULATED VESSEL APPROACHES 649<br />

WILLIAMS,R.,D. E. BAIN,J.K.B.FORD AND A. W. TRITES. 2002a. Behavioural responses of<br />

male killer whales to leapfrogging vessel. Journal of Cetacean Research and Management<br />

4:305<strong>–</strong>310.<br />

WILLIAMS, R.,A.W. TRITES AND D. E. BAIN. 2002b. Behavioural responses of killer whales<br />

(Orcinus orca) to whale-watching boats: Opportunistic observations and experimental<br />

approaches. Journal of Zoology 256:255<strong>–</strong>270.<br />

Received: 9 May 2006<br />

Accepted: 5 February 2007


MARINE MAMMAL SCIENCE, 15(1): 102-122 (January 1999)<br />

0 1999 by the Society for Marine Mammalogy<br />

BEHAVIORAL SAMPLING METHODS FOR<br />

CETACEANS: A REVIEW AND CRITIQUE<br />

JANET MANN<br />

Department of Psychology and Department of Biology<br />

Georgetown University<br />

Washington, DC 20057, U.S.A.<br />

E-mail: mannj2@gunet.georgetown.edu<br />

ABSTRACT<br />

Behavioral scientists have developed methods for sampling behavior in or-<br />

der to reduce observational biases and to facilitate comparisons between stud-<br />

ies. A review of 74 cetacean behavioral field studies published from 1989 to<br />

1995 in Marine Mammal Science and The Canadian Journal of Zoology suggests<br />

that cetacean researchers have not made optimal use of available methodology.<br />

The survey revealed that a large proportion of studies did not use reliable<br />

sampling methods. Ad libitum sampling was used most often (59%). When<br />

anecdotal studies were excluded, 45% of 53 behavioral studies used ad libitum<br />

as the predominant method. Other sampling methods were continuous, one-<br />

zero, incident, point, sequence, or scan sampling. Recommendations for sam-<br />

pling methods are made, depending on identifiability of animals, group sizes,<br />

dive durations, and change in group membership.<br />

Key words: methods, sampling, observations, behavior, vocalization, Ceracea,<br />

ethology, quantitative.<br />

Researchers studying the behavior of cetaceans at sea face several unusual<br />

methodological challenges. Many cetaceans swim rapidly, range over long dis-<br />

tances on a daily basis, and have seasonal migrations of thousands of kilome-<br />

ters. Cetaceans are difficult to follow because they disappear during dives and<br />

do not leave long-lasting traces, such as tracks, scats, or dens.<br />

To meet these challenges, cetacean researchers have developed photoiden-<br />

tification methods and some clever technical solutions. For example, individ-<br />

uals of many species can be identified by natural markings, allowing research-<br />

ers to link sightings separated by years and thousands of kilometers (Ham-<br />

mond et al. 1990). Individual animals can be followed in more detail by<br />

tagging them with radio, acoustic, or satellite tags. Diving behavior can be<br />

monitored using time-depth recorders. Hydrophones and underwater video<br />

cameras can provide continuous data on the behavior and communication sig-<br />

nals of animals for shorter periods of time.<br />

In spite of these impressive technical advances, methods for sampling ce-<br />

102


MANN: BEHAVIORAL SAMPLING METHODS 103<br />

tacean behavior have received less attention. In this paper, I review quantitative<br />

observational methods typically used in cetacean studies and offer specific sam-<br />

pling suggestions.<br />

Altmann (1974) points out that most field observations of behavior involve<br />

sampling decisions, whether those decisions are explicit or not.<br />

“Sampling decisions are made wbenever the stua’ent of social behavior cannot continuously<br />

observe and record all of the behavior of all of the members of a social group. . , . We<br />

suspect that the investigator ofen chooses a sampling procedure without being aware that<br />

he is making a choice. Of course, he does not thereby escape fiom the consequences of that<br />

choice.” (Altmann 1974:229).<br />

The lack of an explicit protocol for sampling was termed ad libitum sam-<br />

pling by Altmann (1974), and she pointed out that it typically entails scoring<br />

“as much as one can” or whatever is most readily observable of the behavior<br />

of a group (Altmann 1974:235-236). These kinds of observations may be<br />

necessary as one learns to identify behaviors, as one plans a study, or as one<br />

observes rare but significant events. However, ad libitum observations suffer<br />

from a variety of potential biases. Different individuals may be more or less<br />

visible. Some behaviors may be more salient and more readily recorded than<br />

others. Individual animals may alter their behavior depending on how visible<br />

they are to other animals (and to observers). The same observer may concen-<br />

trate on different behaviors during different observation periods, and different<br />

observers may notice or attend to entirely different behaviors during the same<br />

observation period. Such biases indicate that ad libitum data are not appropriate<br />

for estimating rates of behaviors or for comparing rates across subjects or across<br />

studies. The selection and appropriate use of sampling methods that yield<br />

unbiased estimates of behavior are critical to the scientific validity of any study.<br />

In this paper, first I review the methods employed in the majority of recent<br />

cetacean field studies, including sampling method and protocols, and discuss<br />

the general advantages and pitfalls of each sampling method. Second, I rec-<br />

ommend specific sampling methods for cetaceans, depending on identifiability<br />

of animals, group sizes, dive durations, and change in group membership.<br />

Literature Survey<br />

To evaluate what methods are currently used by cetologists, I surveyed<br />

papers on cetacean behavior published from 1989 through 1995 (see Appendix<br />

1 for details). To limit the survey, I selected two peer-reviewed journals that<br />

together published the majority of studies in wild cetacean behavior, The Ca-<br />

nadian Journal of Zoology (CJZ, 31% of studies) and Marine Mammal Science<br />

(MMS, 29% of studies).<br />

Seventy-four studies were reviewed, 38 in CJZ and 36 in MMS. The fol-<br />

lowing information was noted: species, age and sex classes, number of animals,<br />

number of observation or recording hours, group-size minima and maxima,<br />

whether animals were individually identified, types of behaviors recorded (in-<br />

cluding vocalizations), definition of group and definition of behaviors, and


104 MARINE MAMMAL SCIENCE, VOL. 15. NO. 1. 1999<br />

methods of observation and sampling. In many cases, essential information in<br />

these categories was not provided (Appendix 1).<br />

CRITIQUE OF METHODS<br />

For every behavioral study, two basic kinds of sampling decisions must be<br />

made. One choice concerns which subject(s) one watches and for how long;<br />

the other concerns the details of how behavior is recorded (Martin and Bateson<br />

1986). I will discriminate these as “follow protocol” and the “sampling meth-<br />

od.” “Follow protocol” refers to how long an observation extends and to wheth-<br />

er researchers follow a group or an individual animal. “Sampling method”<br />

refers to the procedures used to sample the behavior of individuals or groups.<br />

Such a distinction is necessary because the sources of error or bias differ ac-<br />

cording to each protocol and according to each method. For example, in the<br />

ethological literature the term “focal-animal sampling” is used to refer to data<br />

collection that involves sampling of behavior of one individual for a set period<br />

of time. However, there are really two separate components, the follow protocol<br />

(following an individual) and a sampling method (which could be continuous<br />

sampling, point sampling, etc.). To state only that one’s method is “focal-<br />

animal sampling” would be insufficient. For example, a researcher could collect<br />

data on a focal individual systematically at regular intervals or opportunisti-<br />

cally (ad libitum ) by irregularly noting behaviors of interest.<br />

In the following sections, the prevalence and the costs and benefits of dif-<br />

ferent follow protocols and sampling methods are discussed. Table 1 is a guide,<br />

illustrating how different techniques (follow protocol and sampling method)<br />

are likely or unlikely to be effective, depending on the characteristics of the<br />

animals under study. My use of the term “identifiable” in Table 1 refers to<br />

cases when observers can identify known individuals (such as in a longitudinal<br />

study) or can discriminate between animals sufficiently to keep track of the<br />

same animal. When I refer to “group,” both for group size and group mem-<br />

bership, this concerns the number of animals close enough together to be<br />

potentially confused with each other.<br />

Follow Protocol<br />

For the studies surveyed, five different follow protocols can be defined:<br />

survey, group-follow, individual-follow, tracking, and anecdote. Three studies used<br />

more than one protocol (e.g., group-follows and surveys). Two studies did not<br />

provide enough information to allow classification by protocol.<br />

Survey refers to encountering groups or individual animals and staying with<br />

those animals for brief periods to census, for example, the number of animals,<br />

identifications, location, and behavior. If observers typically monitor groups<br />

for 30 min or less, then their protocol is identified here as “survey.” Surveys<br />

comprised 16% (n = 12) of the studies reviewed. Surveys provide a “snapshot”<br />

of animal life and are very useful for tracking patterns of association and for<br />

analyzing demographic, reproductive, and ecological factors. Surveys are par-


MA”: BEHAVIORAL SAMPLING METHODS 105<br />

Table 1. Recommended uses for different sampling methods depending on subject<br />

characteristics.<br />

Group<br />

Rapidly Group membership Dive time<br />

identifiable? size rate of change<br />

Method yes no*


106 MARINE MAMMAL SCIENCE, VOL. 15, NO. 1, 1999<br />

Appropriate sampling techniques applicable during group-follows are dis-<br />

cussed below (see Table 1).<br />

If observers monitor an individual regardless of whether the animal is in a<br />

group or not, their protocol is classified as indiuidaal-follow. Only 12% of the<br />

studies reviewed used this protocol. The indiuidad-follow is roughly equivalent<br />

to focal-animal sampling. The critical feature of this method is to focus on<br />

one animal and systematically record behavior that is defined a priori. When<br />

sampling individual behavior, researchers may still collect ad libitam data on<br />

other animals or scan the group at regular intervals to determine group mem-<br />

bership. However, with few exceptions, data collection on non-focal animals<br />

should not compromise focal data.<br />

The individual-follow enables the observer to focus on the individual ani-<br />

mal’s “perspective.” What is the day in the life of that animal like? Who do<br />

they approach, avoid, stay close to, interact with, mate with, and fight with?<br />

Observing the continuous stream of individual behavior in different contexts<br />

is central to the understanding of the dynamics of social relationships. Time-<br />

budget data for individual animals are more appropriate for most studies than<br />

group-activity data because individuals of varying age and sex have different<br />

behavioral and ecological strategies. Both methodological and theoretical ar-<br />

guments support focusing on the individual as the unit of analysis. Selection<br />

operates at the level of the individual (Williams 1966), and this perspective<br />

is critical to understanding the behavioral and reproductive strategies of ce-<br />

taceans. Group-living cetaceans may rely on each other for survival and repro-<br />

duction, but the costs and benefits of group living are unlikely to be shared<br />

equally among all group members. Measures of such variation are particularly<br />

relevant to the study of how natural and sexual selection has shaped the evo-<br />

lution of social systems and behavior patterns in a species.<br />

The success of systematic sampling with the individual-follow depends on<br />

how rapidly animals can be identified or discriminated from other group mem-<br />

bers when they surface (Table 1). Several factors influence this, including the<br />

size and stability of groups and the duration of dives or “out-of-sight’’ periods.<br />

The optimal situation for individual-follows is that of a small stable group<br />

where individuals can be identified as soon as they come into view. If indi-<br />

viduals are not readily identifiable upon surfacing, individual-follows may still<br />

be used if the animals are solitary much of the time (e.g., humpback whales).<br />

Group size is also a factor. No matter how identifiable an animal is, if there<br />

are hundreds of animals, it will be hard to keep track of a focal animal (Table<br />

1). Brief focal samples, spanning 5 min or less (one or two surfacing bouts),<br />

may. be appropriate under such conditions. Individual-follows might be facil-<br />

itated by focusing on a readily identifiable member (i.e., based on size, marks,<br />

tags, or radiotracking), although the potential biases inherent to such selection<br />

must be considered.<br />

Although sampling decisions depend on the observation conditions, long<br />

individual focal-follows are highly recommended where possible. Rogosa and<br />

Ghandour (1991) have argued (mathematically) that short samples are one of<br />

the biggest sources of unreliability in sampling uncommon behaviors. Long


MANN: BEHAVIORAL SAMPLING METHODS 107<br />

individual follows may also be practical, given the search-time costs of sam-<br />

pling different individuals on the same day. If the researcher is likely to lose<br />

track of the focal animal in certain behavioral contexts (e.g., foraging), then<br />

scans or other sampling techniques may be applied to prorate or correct for<br />

such biases. Individual follows often enable the observer to identify the sources<br />

of sampling bias (e.g., conditions that prove difficult for monitoring individual<br />

behavior). If the study design requires following individuals for long periods<br />

of time, but individuals are not rapidly identifiable, then visual, radio, or<br />

acoustic tags may be necessary.<br />

Tracking refers to studies that electronically monitor individuals’ locations<br />

or behavior (through hydrophones, transponder tags, or other devices). Track-<br />

ing is particularly valuable if researchers need to continuously record the be-<br />

havior of an animal over long periods. Six studies (8%) reviewed here tracked<br />

small numbers of animals using radio and transponder tags to record diving<br />

and vocal behavior. This approach can be relatively expensive, and the attach-<br />

ment process or device may affect behavior. Sample sizes tend to be small in<br />

tracking studies.<br />

An anecdote is a descriptive report of a single event or series of events, such<br />

as birth or predation. Twenty-eight percent (n = 21) of the studies analyzed<br />

here were classified as anecdotes. Anecdotes are a valuable means of describing<br />

rarely observed events.<br />

The survey and group-follow may be considered “group-protocols”-where<br />

groups are monitored unless animals are alone. The individual-follow is an<br />

“individual-protocol”-where a specific individual is monitored regardless of<br />

whether or not it is a member of a group. Tracking can involve groups or<br />

individuals. The merits of different protocols must be weighed against their<br />

impact or influence on the animal’s behavior. Few data are available on whether<br />

or how tagging, equipment noise (e.g., engine, tag, depth sounder), and pro-<br />

longed vs. brief follows directly affect the behavior of cetaceans. Indirect effects<br />

may also occur by alteration aQ the behavior of cetacean prey or predators. By<br />

combining different approachies, or directly testing for the effects of these<br />

approaches, observers can find ways to minimize their impact on behavior.<br />

Sampling Methods<br />

Given that a researcher has, chosen a protocol for following the animals, a<br />

number of sampling methods may be used. These include ad libitum, contin-<br />

uous, fical-group, one-zero, point, scan, predominant activity, sequence, and all-event1<br />

incident sampling (defined below). The success of each sampling method de-<br />

pends, in part, on whether a researcher is focusing on events (brief behaviors,<br />

measured in frequency) or states (long behaviors of measurable duration). Al-<br />

though the distinction is convenient, events and states are on a continuum<br />

(ie., all events have durations and all states have frequencies).<br />

Ad libitum sampling-Ad libitum samples are “typical field notes” (Altmann<br />

1974). The observer writes down what seems of interest. Ad libitum sampling<br />

does not involve systematic constraints on what is recorded and when it is


108 MARINE MAMMAL SCIENCE, VOL. 15, NO. 1, 1999<br />

recorded. The initial phases of behavioral study often involve some ad libitzlm<br />

sampling in order to delineate and define behaviors and research questions.<br />

Most observers continue to collect some ad libitum data throughout the course<br />

of their study. In the survey of studies, the sampling method was classified as<br />

ad libitum only if it was the predominant method used. Ad libitum sampling<br />

was the most common technique used (59% of studies). Ad libitum sampling<br />

is useful for certain kinds of comparisons, but not for estimating rates of<br />

behavior or for comparing behavior patterns of different age or sex classes<br />

(Altmann 1974). As Altmann points out, some animals may be more notice-<br />

able, either because of their behavior (or activity level), size, or level of habit-<br />

uation. Ad libitum sampling may be applied during individual focal-follows<br />

(e.g., prey captures by focal animal) or group-follows (e.g., displays). Ad libitum<br />

data are valuable for looking at the direction of interactions or at what happens<br />

within a dyad if the direction of the interaction is unlikely to be influenced<br />

by the probability of seeing it. For example, ad libitum data are commonly<br />

used in ethology to assess dominance relationships, where one animal “wins,”<br />

and the other “loses” an agonistic interaction (e.g., see Samuels and Gifford<br />

1997). Rates of agonistic interactions cannot be acquired through ad libitum<br />

sampling, because the likelihood of seeing an interaction may be affected by<br />

the rank, size, sex, or other characteristics of the animals (e.g., conflicts between<br />

adult males are more likely to escalate and thus be observed than conflicts<br />

between adult females). However, given that two animals are fighting, the<br />

outcome of the fight is unlikely to influence or be influenced by the probability<br />

of seeing the interaction. Ad libitum sampling is effective for examining di-<br />

rected displays, in which both signaller(s) and recipient(s) can be determined.<br />

Sociomatrices are an excellent means of illustrating some patterns of interac-<br />

tion with ad libitum data. A sociomatrix has all potential interactants listed<br />

on the top and along the side, with one axis indicating actor or “winner” and<br />

the other as recipient or “loser.” Frequencies are tallied within each cell of<br />

interactants. When dominance hierarchies are linear, for example, then most<br />

entries will fall on one side of the diagonal. Sociomatrices can be applied to<br />

other behaviors sampled ad libitum. Mann and Smuts (in press) recorded 1,3 1 1<br />

rubbing events between nine wild Tursiops newborns and their mothers during<br />

focal individual-follows (189 h). Rubbing rate could not be determined, nor<br />

whether some infants rubbed more than others. But rubbings were extremely<br />

asymmetrical, with infants initiating and performing 99% of the rubbing on<br />

their mothers.<br />

Ad libitum data are a valuable part of any field study but should not be<br />

represented as rates, proportions, frequencies, or other unbiased estimates of<br />

behavior. The use of ad libitum data for anecdotes is perfectly appropriate. Rare<br />

events such as predation, a birth, or a lethal fight, can provide critical infor-<br />

mation and insights, and ad libitum sampling is likely to be the only way such<br />

events are recorded.<br />

Continuous sampling-Alternate names: event sampling, frequency sampling,<br />

focal-animal sampling. Continuous sampling is a systematic record of fre-<br />

quencies or durations for a specified set of behaviors. Researchers occasionally


MANN: BEHAVIORAL SAMPLING METHODS 109<br />

use the term “focal-animal sampling,” to indicate continuous sampling of<br />

behavior during an individual-follow. Measuring the exact times (and dura-<br />

tions for behavioral states) of every occurrence of a behavior is very demanding<br />

for the observer. Scoring event frequencies within time blocks can simplify<br />

data collection. The reliability of continuous data can be easily compromised<br />

if the observer tries to record too many behaviors at once, especially when the<br />

animals are active. Thus, Altimann (1974) recommends that this method be<br />

used for one or two animals at most.<br />

Continuous data on associa.tions, diving behavior (with transponder tags),<br />

other behaviors, or vocalizations were collected in 14% (n = 10) of the studies<br />

surveyed. In four of these studies researchers observed animals directly. Two<br />

studies used tags to track individuals. The remaining four collected continuous<br />

data on small groups of animals using videotape or multiple observers.<br />

Continuous sampling of behavior is relatively simple for activities at the<br />

surface (during surfacing bouts). Surfacing-bout durations, breath frequency,<br />

dive types, surface-display rates, and synchronous surfacings of the focal animal<br />

and others may be recorded on a continuous basis for many dolphin and whale<br />

species. Even deep-diving animals, such as sperm whales, may have prolonged<br />

bouts of socializing at the surface (Whitehead and Weilgart 1991). If the<br />

activities are brief, or difficult to time, then they can be scored as events<br />

(frequencies), rather than states (onset to offset). Records of behavior sequences<br />

are typically intact in continuous sampling. As Table 1 indicates, continuous<br />

sampling is most likely to succeed when animals are rapidly identifiable, live<br />

in small groups, and dive for short periods. Under these sampling conditions,<br />

observers can keep track of a specific individual for long periods of time.<br />

Continuous data are the richest source of information on social behavior and<br />

relationships, because such dlata include information on details, sequences,<br />

actors and recipients, rates and durations of behavior for individual animals.<br />

While it is simple to record vocalizations continuously using a tape recorder,<br />

for these data to qualify as continuous data for an individual follow, the ob-<br />

server must be able to identify which vocalization comes from the focal in-<br />

dividual. Use of tags that can record or transmit acoustic data is one technique<br />

for achieving this (e.g,, Tyack 1985, Tyack and Recchia 1991). Another tech-<br />

nique involves passive acoustic localization (e.g. Clark 1980, Freitag and Tyack<br />

1993, Frankel et al. 1995). Fusion of these acoustic and visual data remains a<br />

challenge to anyone interested in cetacean communication. As with behavioral<br />

studies, the study of cetacean vocalizations still hinges upon such basic infor-<br />

mation as the species, sex, or age class of the animal producing the call. The<br />

study of how these sounds are used in social communication will require<br />

increased application of methods to identify which animal produces which<br />

sound during interactions wit:hin and between groups.<br />

Focal group sampling-Alternate names: focal subgroup sampling, group<br />

sampling, predominant group activity sampling. Focal-group sampling is a<br />

continuous assessment of group activity. Group activity may be scored at in-<br />

tervals (e.g., indicate predominant group activity every 5 min) or continuously<br />

(e.g., the group rested from 1122-1147). This method is widely used in ce-


110 MARINE MAMMAL SCIENCE, VOL. 15, NO. 1, 1999<br />

tacean research when observers follow groups, and Altmann (1974) has been<br />

cited in support of it (e.g., Shane 1990). Altmann (1974) used the term “focal<br />

sub-group sampling” to refer to continuous sampling with more than one<br />

individual. However, Altmann described focal sub-group sampling as appro-<br />

priate only under a very restricted set of conditions “in which all individuals<br />

in the sample group are continuously visible throughout the sample period<br />

. . . if one is working with observational conditions that are less than perfect,<br />

focal-animal sampling should be done on just one focal individual at a time,<br />

or at most, a pair (e.g., mother and young infant)” (Altmann 1974:243-244).<br />

Twenty-seven percent of the studies surveyed here (45% of ad libitum studies)<br />

used “focal group sampling.” These studies scored “group behavior.” Dozens<br />

of animals, not pairs, were sampled using focal-group sampling. I chose to<br />

classify these studies as ad libitum because none of the conditions necessary for<br />

unbiased focal-group-sampling were met. That is, observers could not have<br />

continuously observed all animals equally in a group regardless of activity.<br />

Some observers appeared to visually assess group activity, perhaps by infor-<br />

mally scanning the group. However, this method is not explicit enough in<br />

how subjects were sampled and does not differ in any substantial way from<br />

ad libitum sampling.<br />

A few studies identified their method as predominant group-activity sam-<br />

pling. This should not be confused with predominant activity sampling (de-<br />

fined below). Predominant group-activity sampling is essentially the same as<br />

“focal group sampling” and was coded as ad libitum in the survey. During<br />

predominant group-activity sampling, the observer defines group activity<br />

based on an assessment of what most (>50%) of the group is engaged in over<br />

an interval, focusing on the proportion of individuals estimated to be engaging<br />

in a behavior, not the proportion of time an individual engages in a behavior for<br />

that interval. An estimate of predominant group activity can be achieved by<br />

explicitly scan sampling over 50% of the individuals in the group, rather than<br />

by “watching the group.”<br />

Altmann (1974) prescribed a restricted set of conditions for focal sub-group<br />

sampling because of the many biases inherent to watching groups. First and<br />

foremost, the observer’s attention is naturally drawn to animals that are most<br />

visible, most active, or most interesting. No observer can possibly continuously<br />

track and record all behavior of all individuals simultaneously under all con-<br />

ditions; it is difficult enough to record the behavior of one animal. Group<br />

activity may be determined if all members of a small group are cohesive and<br />

engaging in the same behavior, such as resting closely at the surface, but it<br />

may be difficult to determine if some animals are engaging in other activities.<br />

The accuracy of focal-group sampling is dependent upon group size, cohe-<br />

siveness, and activities of the animals, thus potentially introducing biases into<br />

data collection. Some behaviors, such as socializing, may be particularly ob-<br />

vious to observers, and even though most of the animals are resting, one might<br />

score the more visible behaviors as group activity. While using group-sam-<br />

pling, observers are often making implicit assumptions about the relative im-


MA”: BEHAVIORAL SAMPLING METHODS 111<br />

portance of behaviors of different age and sex classes. If two mothers are hunt-<br />

ing, but their calves are traveling, should one call it hunting or traveling?<br />

With video, continuous sampling of groups is possible because observers<br />

can rewind the tape and code the behavior of different individuals separately<br />

(thus making the process equivalent to conducting multiple continuous focal-<br />

animal samples simultaneously). However, simultaneous behaviors of different<br />

individuals within the same group are likely to depend upon one another,<br />

which must be taken into account for statistical analyses. Focal-group sam-<br />

pling is not only used during the group-follow protocol, it is also used during<br />

surveys to identify “group activity.” However, this too is subject to the same<br />

criticisms.<br />

One study (Fragaszy et a/. 1992) has compared focal-group sampling with<br />

focal-individual point sampling and scan sampling (see definitions below).<br />

This study, although on primates, is particularly relevant for cetacean studies<br />

because, like cetaceans, capuchins and squirrel monkeys move in and out of<br />

view and are difficult to follow for continuous periods of time. Fragaszy et al.<br />

(1992) compared foraging time budgets of squirrel monkeys using both focal-<br />

group sampling and focal-animal point sampling methods and concluded that<br />

focal-group sampling overestimated foraging considerably. They came to this<br />

conclusion by correlating individual-focal point sampling rates with focal-<br />

group sampling rates. The focal-group sampling error rate (computed by JM)<br />

ranged between 39% and 63%. They also compared scan sampling of capuchin<br />

monkey groups to focal-animal point sampling. Scan sampling fared much<br />

better. When time budgets for scan sampling and focal-animal point sampling<br />

for foraging among capuchins were compared, the error rate was only 0.8%<br />

(computed by JM). For other, less frequent behavioral states, the error rates<br />

for scan sampling ranged from 10% to 26%. Focal-group sampling error rates<br />

for other behavioral states were not compared. However, error rates are lowest<br />

for more prevalent behaviors, and foraging was the most common state. Thus,<br />

error rates are likely to increase for other states. Such error rates should give<br />

pause to anyone considering group-sampling.<br />

One-zero sumpling-Alternate names: time-sampling, Hansen frequencies, in-<br />

terval sampling, partial interval sampling, method of repeated short samples,<br />

and modified frequencies. One-zero sampling entails scoring whether or not a<br />

behavior occurs during an interval (e.g., 30 sec), rather than scoring how fre-<br />

quently or how long the behavior occurred. For example, an observer may<br />

score whether or not an animal vocalized during 30 seconds, or whether or<br />

not it rubbed, rather than the frequency of vocalization events or the duration<br />

of rubbing (a state). Nine percent of the studies surveyed employed one-zero<br />

sampling. With few exceptions, this technique is not recommended, because<br />

one-zero scores do not represent frequency or duration and are prone to very<br />

high rates of sampling error (see Mann et al. 1991). Although one-zero sam-<br />

pling has no recommended uses (Altmann 1974), there are some natural types<br />

of one-zero data (categorical variables) that are not represented by the flow of<br />

continuous behavior but may occur over very long intervals. For example, if<br />

group fissions and fusions are rare, then whether or not an animal joins a


112 MARINE MAMMAL SCIENCE, VOL. 15, NO. 1, 1999<br />

group during a day or field season is a significant piece of information and<br />

may be more biologically meaningful than the absolute rate. However, one<br />

must already know something about the rates and durations of behaviors of<br />

interest before employing one-zero sampling.<br />

Researchers sometimes use one-zero events to define states. That is, one-<br />

zero data may assist in the development of an ethogram but not be part of<br />

the sampling method. For example, if two animals are seen in contact during<br />

an interval, this may “define” socializing regardless of the duration or fre-<br />

quency of contact. Or, one might define foraging based on whether or not the<br />

animal flukes-up at the dive. It is important then to distinguish between the<br />

use of one-zero (categorical) information to help define behaviors that are pre-<br />

sumed to be continuous and treating one-zero scores as state behaviors them-<br />

selves.<br />

Point sampling-Alternate names: instantaneous sampling, on-the-dot sam-<br />

pling, time sampling. Point sampling entails scoring activity as a “snapshot”<br />

at a given moment (e.g., every 30 sec). The observer may score states, such as<br />

distance to others, activity, or other information on a point-sampling basis.<br />

Five percent of the studies surveyed used point sampling. It is a reliable<br />

method that is widely applied in ethological studies but has been rarely used<br />

for cetaceans. One possible explanation is that the “point” often happens when<br />

the animal is submerged. The few studies reviewed here that used point sam-<br />

pling did so by electronically tracking the animal. For direct observations,<br />

point sampling can still be applied with two basic strategies. (1) The animal’s<br />

behavior is sampled at the first surfacing after the interval point. Two factors<br />

may lead to sampling biases. One ends up sampling surface behavior rather<br />

than the behavior occurring underwater at the point. By examining the rela-<br />

tionships between surface and subsurface activities, these biases may be min-<br />

imized. Another bias is that the behavior is not sampled at regular intervals,<br />

but at intervals determined in part by the subject’s surfacing and diving be-<br />

havior. If the observer sampled at 1-min intervals, and animals tended to<br />

remain close to the surface during socializing and dive deeply for longer pe-<br />

riods during foraging, then the observer could successfully sample socializing<br />

but would miss several sampling intervals during foraging. Social behavior<br />

would be overrepresented in relation to foraging. This problem can be alle-<br />

viated by establishing a point interval greater than the typical long dives of<br />

the subject. (2) The second strategy is to use point sampling for behavioral<br />

states if one can assume that the state continues when the focal animal moves in and<br />

out of view. Observers often infer behavioral state based on surfacing pattern,<br />

proximity of other individuals, and other cues (e.g., flukeprints or bubbles at<br />

the surface). The observer may apply a convention such as scoring a behavioral<br />

state at the “point” if the animal engaged in the behavior prior to diving or<br />

disappearing from view and was engaged in the same behavior upon surfacing.<br />

This convention may be applied if the “out-of-sight” periods tend to be much<br />

shorter than the bout duration of the behavioral state. Another convention is<br />

to score the behavior as that last observed, or, alternatively, first seen upon<br />

surfacing. Such conventions must be explicitly defined, justified, and the po-


MANN: BEHAVIORAL SAMPLING METHODS 113<br />

tential biases considered. As Table 1 suggests, point sampling can be applied<br />

under many conditions but becomes problematic if dive times or out-of-sight<br />

periods are long and group sizes are large (making it difficult to keep track<br />

of an individual animal).<br />

Point sampling can be a very useful method to determine time budgets or<br />

diurnal patterns of behavior. This method is also a useful supplement to other<br />

focal-individual sampling methods. For example, the observer may record an<br />

animal’s speed at 5-min intervals, or note a focal animal’s nearest neighbor<br />

based on who surfaces closest to the focal animal at the first surfacing after<br />

each 10-min interval. Although point sampling is typically recommended for<br />

behaviors of appreciable duration, activity can be more difficult to score on a<br />

point-sampling basis. In reality, even with the animal in full view, it may take<br />

several seconds or more to “decide” what an animal is doing. It is important<br />

that the observer decide as quickly as possible, so shelhe does not inadvertently<br />

wait until the animal does something that is easier or more interesting to<br />

score. Brief behaviors or events are difficult to observe accurately with point<br />

sampling because such behaviors are often missed at the “point.”<br />

Scan sampling-Scan sampling entails taking a “point” or “instantaneous”<br />

sample of an individual’s behavior or location before moving to the next animal<br />

and doing the same. Scans are conducted either at regular intervals (e.g., sample<br />

each animal at 10-sec intervals), or as quickly as possible (ie., search for the<br />

next animal as soon as the last was sampled). Both point and scan samples are<br />

good techniques for measuring states, but brief events are likely to be missed.<br />

Three percent of the cetacean field studies reviewed here used scan sampling.<br />

Scan sampling is very valuable for sampling behavior when focal-individual<br />

observations are not possible or desirable, or if the researcher wishes to keep<br />

track of group activities. An explicit scan-sampling technique should also be<br />

considered for surveys of groups, rather than using ad libitam sampling to<br />

identify “group activity.” Observers sometimes use a random scan-sampling<br />

schedule, or just scan from one side of a group to the other. Alternatively, if<br />

group sizes tend to be large and the animals move rapidly, but age or sex<br />

classes are distinctive in size or markings, then the researcher may scan age<br />

and sex classes separately (e.g., scan large males, then females with dependent<br />

calves, then immature animals). This would leave some agehex classes unsam-<br />

pled but would enable the observer to scan a group of fast-moving animals<br />

more effectively.<br />

Scan sampling and point sampling share practical difficulties: the time it<br />

takes to decide the animal’s activity and out-of-sight or diving periods. One<br />

may employ the same strategies for scan sampling as for point sampling, but<br />

with scan sampling the observer samples each individual only once per session.<br />

In some cases, the observer may need to watch each individual for 3 min or<br />

more (approximating a short individual-follow) but uses the midpoint (e.g.,<br />

minute 2) to define the behavioral state. Thus, one data point is taken for<br />

each animal in succession, although each observation period is longer than for<br />

a typical scan. The same conventions used for point-sampling out-of-sight<br />

periods can be adapted to scan sampling.


114 MARINE MAMMAL SCIENCE. VOL. 15. NO. 1. 1999<br />

Potential uses for scan sampling are varied (see Table 1). First, an observer<br />

may scan a group at fixed intervals (e.g., every 10 min) to assess which is still<br />

present, which is close to a focal animal, etc. Second, an observer may scan a<br />

group to rapidly estimate group activity by identifying each animal’s activity<br />

as it surfaces. This is only possible for groups where one can keep track of<br />

each individual within the group during surfacing and depends in part on the<br />

degree of interindividual overlap in surfacing bouts. The observer may be able<br />

to scan rapidly enough to minimize the likelihood of resampling the same<br />

individual or scan from the front to the back of a group if direction changes<br />

are infrequent. If the scan procedure inadvertently allows for repeat sampling<br />

within a session, then the resampling of highly identifiable individuals would<br />

alert the observer to this problem. Third, an observer might use it to assess<br />

nearest neighbors for each animal. Fourth, one could record both activity and<br />

nearest neighbor for each group member. If group scans are conducted during<br />

individual-follows, then it is important that the scans can be completed very<br />

quickly to prevent the observer from being distracted from the focal individ-<br />

ual. With scan sampling, cetacean researchers can broaden their dataset to look<br />

at coordination of group activities and refined measures of association for a<br />

number of animals of different age and sex classes.<br />

Predominant activity sampling-Predominant Activity Sampling (PAS), de-<br />

veloped by Hutt and Hutt (1970), refers to scoring individual behavior as the<br />

predominant activity over some interval (e.g., 30 sec), only if that behavior<br />

occupied 50% or more of that interval (>15 sec). This method is only useful<br />

for measuring states. No studies in the survey used PAS. It is an empirically<br />

valid technique for estimating the proportion of time during which behavioral<br />

states occur (Tyler 1979). Very brief behaviors or displays (events) will not be<br />

represented in PAS data unless the observer uses very short intervals. It is<br />

important that the interval length is brief enough to capture the briefest states<br />

of interest.<br />

PAS is useful for sampling animals that go in and out of view for brief<br />

periods. In a sample of Ti~rsiops calves, some types of data have been collected<br />

using both focal individual point samples and focal PAS simultaneously (J.<br />

Mann, unpublished data). Thus calf activity budgets could be compared based<br />

on sampling type. For example, “calf position swimming” (when the calf is<br />

in contact under the mother) was measured using both point sampling (in-<br />

stantaneous measures of swim position at first surface after each 2.5-min in-<br />

terval) and PAS (calculated from the continuous data on observed onsets and<br />

terminations of calf position swimming). I compared the percent time swim-<br />

ming in calf position for 19 calves in my longitudinal sample. Each calf was<br />

observed during focal-individual follows for 10-15 h per yr for one or more<br />

years. The sum of observation hours for all calves across all years was 750.<br />

Each infant had a PAS and point sampling percentage for each year, for a total<br />

of 40 calf-years (an average of two years or samples per calf). The PAS per-<br />

centages and point sampling percentages were correlated by treating each in-<br />

fant observation year as independent. This yielded a significant value (Spear-<br />

man’s rho = 0.97, n = 40, P < 0.0001). The mean percent error rate for PAS


MA”: BEHAVIORAL SAMPLING METHODS 115<br />

was 3.2%. Several methodologists (e,g., Tyler 1979) have demonstrated the<br />

validity of PAS and point sampling using statistical models, but no one has<br />

contrasted these methods using actual observational data.<br />

Incident sampling-Alternate names: all-event sampling, all-occurence sam-<br />

pling, all-animals sampling. Incident sampling entails scoring all behavioral<br />

events of a specific type in a group. This method is not applicable for most<br />

behavioral states. The observability of the events is key to the success of this<br />

method. The behaviors themselves must be obvious enough to alert the ob-<br />

server (e.g., breaching). In addition, the observer must be able to record all<br />

the events regardless of how many animals are present. Thus, for incident<br />

sampling to be successful, the behavior must be sufficiently infrequent to allow<br />

complete recording for the group. As Altmann points out, this is a form of<br />

continuous sampling (referred to as focal animal sampling in her paper) of a<br />

group for a restricted set of behaviors. However, with continuous sampling,<br />

the individual animals are identified. During incident sampling, distinctions<br />

between subjects may or may not be made. Observers should still make every<br />

effort to determine whether the same or different animals are exhibiting the<br />

behavior(s). Two areas can be problematic. First, animals may tend to repeat<br />

displays, thus becoming overrepresented in the dataset. Second, if group com-<br />

position changes often (Table 1) it may be difficult to calculate event rate as<br />

a function of group size or structure.<br />

In the survey, 16% of the studies used incident sampling. Studies that<br />

identified their method as “focal-group-sampling” were sometimes coded as<br />

using incident sampling if they scored a very restricted set of obvious behav-<br />

iors. For example, breaches and lobtails are so visible that observers are likely<br />

to detect each occurrence within a group. Similarly, if observers could deter-<br />

mine how many animals were within acoustic range, then incident sampling<br />

could apply to studies of vocalizations.<br />

Incident or all-event sampling is valuable for cetacean researchers who wish<br />

to focus on specific dramatic or easily recognized events that involve few an-<br />

imals. For example, observers could use this method to compare successful and<br />

unsuccessful killer whale hunting attempts via beaching. Dramatic surface<br />

percussive displays, such as breaching or lobtailing, can also be observed re-<br />

liably in groups (Waters and Whitehead 1990). Incident sampling requires<br />

that the observer systematically record every event. Incident sampling may be<br />

adopted with a group protocol, but the observer should still be able to dis-<br />

tinguish which animal (or agehex class member) is engaging in the behavior<br />

to avoid misattributing the behavior pattern equally to all individuals.<br />

Sequence sampling-In sequence sampling, the observer focuses on sequences<br />

of behavior or on particular interactions, rather than individuals, and system-<br />

atically records all relevant behaviors that occur during the event(s), main-<br />

taining the sequences of behaviors in the record (Altmann 1974). Parameters<br />

defining how an interaction begins and terminates must be specified in se-<br />

quence sampling. For example, observers may score sequences of behaviors<br />

among surface-active groups of humpback whales by defining the beginning<br />

of the sequence as “male humpback whale challenges principal escort by ap-


116 MARINE MAMMAL SCIENCE, VOL. 15, NO. 1, 1999<br />

proaching within 50 m” and terminating the sequence when either the chal-<br />

lenger or the principal escort separates more than 50 m from the female for<br />

30 min. It is distinct from incident sampling because during sequence sam-<br />

pling multiple events may occur in a group, but the observer focuses on the<br />

start of the first event seen and records that particular event or interaction to<br />

completion (even if other group members engage in similar behaviors at the<br />

same time). Only one of the studies surveyed used sequence sampling.<br />

Sequence sampling is recommended for very observable behaviors but can<br />

be applied under a broad range of field conditions (see Table 1). The researcher<br />

must be able to determine when the sequence begins and ends and be able to<br />

discriminate between, although not necessarily identify, the interactants. The<br />

individual animals may be hard to identify if they communicate over long<br />

ranges and move in and out of the observer’s view. For example, if the re-<br />

searcher wants to know whether breaching animals attract or repel others, s/<br />

he can use sequence sampling to test whether animals are likely to approach<br />

or leave the animal who breached (e.g., within 10 min of the first breach).<br />

Sequence sampling is excellent for determining the conditional probabilities<br />

of behavior sequences.<br />

RECOMMENDATIONS<br />

Although tightly focused and informative studies peppered the literature,<br />

the survey of CJZ and MMS revealed two shortcomings in studies of cetacean<br />

behavior. First, a large proportion of studies used ad libitum sampling, which<br />

is replete with bias. Second, there was little consistency in reporting behavioral<br />

data, sampling methods, hours of observation, behavioral and group defini-<br />

tions, and number of subjects. This information is critical to evaluation of the<br />

claims of any study. Furthermore, the comparability of results and hopes of<br />

replication are lost if this information is not provided. Information about<br />

equipment (cameras, film, boat, motor, tape recorders, etc. ) was frequently<br />

reported, but basic methodological information regarding subjects and pro-<br />

tocol was sometimes lacking.<br />

To allow evaluation of statistical power, it is important to indicate the<br />

number of animals observed and how long each animal was observed. Re-<br />

searchers usually reported the number of animals identified in the population<br />

but not the number of animals observed. Sample sizes can be determined only<br />

when animals are observed individually. Similarly, only a few studies treated<br />

each subject, rather than each observation, as independent. As Milinski (1997)<br />

points out, this is a form of pseudoreplication, one of the “deadly sins” in the<br />

study of behavior.<br />

Group size and definition of group are needed, as well as a statement of<br />

protocol used if group composition changed. I recommend proximity-based<br />

measures, because this method is quantifiable and does not rely on behavioral<br />

sampling to determine group membership. “Coordinated-behavior” definitions<br />

make implicit assumptions about proximity, because observers cannot assess<br />

the activities of animals who are kilometers away. Furthermore, individuals at


MANN: BEHAVIORAL SAMPLING METHODS 1<strong>17</strong><br />

close range may be engaged in more than one activity that is shared with<br />

those farther away. An explicit definition of group, and a rationale for the<br />

definition in terms of the study goals, is essential.<br />

Specific details of how behaviors were defined and how behaviors were sam-<br />

pled should be included, using either standardized terms from the ethological<br />

literature or descriptions. Only a few studies of those reviewed used standard<br />

sampling terms described by Altmann (1974) and others. All studies of ce-<br />

tacean behavior should provide the basic methodological information described<br />

in this review. Following are more specific recommendations regarding the<br />

uses of different sampling protocols and techniques, with special consideration<br />

of how to reconcile unbiased observations under ideal conditions with the<br />

imperfect conditions that cetacean biologists face. Additional detailed discus-<br />

sion of methods and statistical comparisons of different observational methods<br />

are available elsewhere (see Altmann 1974, Dunbar 1976, Tyler 1979, Martin<br />

and Bateson 1986, Mann et al. 1991, Rogosa and Ghandour 1991).<br />

Follow Protocol<br />

Sampling biases in group-follows and surveys can be minimized by using<br />

scan sampling, incident sampling, or sequence sampling methods. Monitoring<br />

groups (surveys and group follows) is valuable when the goal is to gain a<br />

“snapshot” of group life through surveys (e.g., habitat use by foraging com-<br />

pared to resting groups) when observation conditions do not permit tracking<br />

of individuals (but tracking of groups is possible), when research questions<br />

focus on the synchrony or coordination of group members, or when the study<br />

focuses on sequences of very observable behaviors. To identify the predominant<br />

group activity during a survey, observers can scan-sample >50% of the group<br />

members. This may be possible for large groups of dolphins or whales that<br />

are visible at the surface for long enough to be scanned. If group sizes are in<br />

the hundreds or thousands, then a protocol for random scan sampling of a<br />

subset of the group may be developed. Sampling biases, such as number of<br />

scans per individual as a function of group size, must be taken into account<br />

during data reduction. Researchers may track a group and record all occur-<br />

rences of certain very obvious behaviors, thus using incident sampling. Se-<br />

quence sampling may also be applied if observers are interested in sequences<br />

of behavior during specific types of interactions, such as cooperative hunting<br />

in killer whales. Focal-group sampling is not recommended, because this<br />

method does not explicitly or systematically sample individuals or behaviors<br />

in groups.<br />

Focal-individual sampling methods with the application of continuous,<br />

point, and/or predominant-activity sampling, are critical tools for insuring<br />

reliable estimates of behavior. The individual is the natural unit of analysis<br />

for behavior. Data can be pooled by individual, and data from different indi-<br />

viduals can often be treated as independent (Machlis et ul. 1985). When ob-<br />

servers cannot identify or discriminate which individual produces a behavior<br />

(e.g,, during scan- or all-event sampling), this makes it difficult to identify


118 MARINE MAMMAL SCIENCE. VOL. 15. NO. 1. 1999<br />

independent units for statistical analysis. This can present problems with the<br />

group-follow protocol.<br />

General Sampling Problems: Deep Divers, Large Groups, and Cowespondence<br />

Between Surface and Subsurface Behavior<br />

The correspondence between surface (observable) and subsurface (often<br />

unobservable) behavior is unknown for most studies of cetaceans, but such<br />

information could help in assessing the biases in relying on surface behavior<br />

alone. For example, some animals forage at depth and not during surfacing<br />

bouts. One option is to define the surfacing breaks between foraging dives as<br />

part of the continuous state of “foraging.” Other behaviors may not change at<br />

the surface (e.g., socializing, resting, traveling). Although it is still valuable<br />

to record surface behavior even though no correspondence to subsurface be-<br />

havior is implied, such sampling decisions should be indicated clearly in the<br />

methods section of the article.<br />

All ethologists face difficult decisions when their subjects disappear from<br />

view. These “out-of-sight’’ periods may be treated differently, depending on<br />

the animal under study. Nesting or burrowing animals may be most likely to<br />

disappear from view when feeding their young or engaging in sexual behavior.<br />

Some cetaceans may disappear for long periods when foraging. Deleting the<br />

time that animals are “out of sight” is desirable when animal visibility is not<br />

determined by or biased by animal activity. Other strategies for using “out-<br />

of-sight’’ periods are needed for cetaceans, because some of their activities co-<br />

vary with diving periods (when animals are most likely to be out of view).<br />

Cetaceans may rest, travel, socialize, or engage in other diverse behaviors at<br />

or near the surface. Certain types of social behavior may occur at depth, but<br />

certainly not all of them do. If the out-of-sight periods tend to be short relative<br />

to the bout lengths of the behavior, then ethologists might designate the<br />

activity for the out-of-sight period as either the last activity seen before the<br />

subject disappeared or the first activity seen when the animal reappears. Al-<br />

ternatively, the observer may consider a state continuous only if the last ac-<br />

tivity seen and the first activity seen after a brief out-of-sight period are the<br />

same. This strategy is not effective if some behaviors, such as foraging or<br />

hunting, consistently occur out of an observer’s view. In such circumstances,<br />

cetacean observers may detect foraging using other cues, including bubble<br />

patterns at the surface, echolocation clicks, or intermittent sightings of fish<br />

catches or chases. Sometimes there are signs at surfacing of what the animal<br />

was doing on the preceding dive. For example, bottom-feeding whales may<br />

surface with mud streaming from the mouth (Wiirsig et al. 1985). Thus,<br />

observers may use fleeting observations of events to define states or indirect<br />

cues (rapid surfacing, changing direction) to define “unobservable” states. One<br />

may also tag an individual to track behaviors at depth for comparison with<br />

cues visible at the surface. By combining these techniques, observers can de-<br />

velop behavioral definitions and systematic protocols to capture the range of<br />

cetacean behaviors.


MANN: BEHAVIORAL SAMPLING METHODS 119<br />

Sampling from large groups of hundreds or thousands (e.g., Scott and Per-<br />

ryman 1991) can prove to be difficult. However, Ostman (1994) has success-<br />

fully completed short focal follows on Stenella longirostris in Hawaii, suggesting<br />

that it is possible to stay with individuals long enough to get a 5-min sample<br />

using an underwater observation booth built into a vessel. Scan sampling of<br />

large groups of unidentified animals is another possible approach. In such<br />

circumstances, researchers would need to scan a subset, possibly by randomly<br />

selecting subgroups or particular agehex classes. Videotaping for later scoring<br />

or photogrammetric analysis (.g., Scott and Perryman 1991) can provide data<br />

on association and behavior.<br />

Although observation conditions for studying cetaceans are rarely ideal,<br />

proximity between animals, synchrony in surfacing, and basic activities can<br />

be recorded systematically under most sampling conditions. Cetacean behav-<br />

ioral research could benefit greatly from a wider use of quantitative sampling<br />

techniques. If applied carefully, such partial but systematic observations can<br />

help elucidate cetacean social and behavioral ecology.<br />

ACKNOWLEDGMENTS<br />

This paper was prepared while I was a fellow at the Center for Advanced Study in<br />

the Behavioral Sciences in Stanford, California, with support from the National Science<br />

Foundation (#SBR-9022 192). Helpful comments on the manuscript were provided by<br />

Robin Baird, Vincent Janik, Edward Miller, Amy Samuels, three anonymous reviewers,<br />

William F. Perrin, Douglas Wartzok, and especially Peter L. Tyack. I was also sup-<br />

ported by the Helen V. Brach Foundation, The Eppley Foundation for Research, and<br />

Georgetown University. I thank Jeanne Altmann for igniting my interest in observa-<br />

tional methods in 1980.<br />

LITERATURE CITED<br />

ALTMANN, J. 1974. Observational study of behavior: Sampling methods. Behaviour 49:<br />

227-267.<br />

CLARK, C. W. 1980. A real-time direction finding device for determining the bearing<br />

to the underwater sounds of southern right whales, (Eubalaena australis). Journal<br />

of the Acoustical Society of America 68:508-5 11.<br />

DUNBAR, R. I. M. 1976. Some aspects of research design and their implications in the<br />

observational study of behaviour. Behaviour 58:79-98.<br />

FRAGASZY, D. M., S. BOINSKI AND J. WHIPPLE. 1992. Behavioral sampling in the field:<br />

Comparison of individual and group sampling methods. American Journal of<br />

Primatology 26:259-275.<br />

FRANKEL, A. S., C.W. CLARK, L.M. HERMAN AND C.M. GABRIELE. 1995. Spatial distribution,<br />

habitat utilization, and social interactions of humpback whales, Megaptera<br />

novaeangliae, off Hawaii determined using acoustic and visual techniques. Canadian<br />

Journal of Zoology 73:1134-1146.<br />

FREITAG, L., AND P. L. TYACK. 1993. Passive acoustic localization of Atlantic bottlenose<br />

dolphin whistles and clicks. Journal of the Acoustical Society of America 93:<br />

2 197-2205.<br />

HAMMOND, P. S. , S. A. MIZROCH AND G.P. DONOVAN. EDS. 1990. Individual recognition<br />

of cetaceans: Use of photo-identification and other techniques to estimate popu-


120 MARINE MAMMAL SCIENCE, VOL. 15, NO. 1, 1999<br />

lation parameters. Report of the <strong>International</strong> Whaling Commission (Special Issue<br />

12).<br />

Hum, S. J., AND C. HUTT. 1970. Direct observation and measurement of behavior.<br />

Charles C. Thomas, Springfield, IL.<br />

MACHLIS, L., P. W. D. DODD AND J. C. FENTRESS. 1985. The pooling fallacy: Problems<br />

arising when individuals contribute more than one observation to the data set.<br />

Zeitschrift fur Tierpsychologie 68:201-2 14.<br />

MANN, J. , AND B. SMUTS. In press. Behavioural development of wild bottlenose dolphin<br />

newborns. Behaviour.<br />

MANN, J., T. TEN HAVE, J. W. PLUNKETT AND S. J. MEISELS. 1991. Time sampling: A<br />

methodological critique. Child Development 62:227-241.<br />

MARTIN, P., AND P. BATESON. 1986. Measuring behaviour: An introductory guide. Cambridge<br />

University Press, Cambridge, UK.<br />

MILINSKI, M. 1997. How to avoid seven deadly sins in the study of behavior. Advances<br />

in the Study of Behavior 26:160-180.<br />

OSTMAN, J. S. 0. 1994. Social organization and social behavior of Hawai’ian spinner<br />

dolphins (Stenella longiristrzs). Ph.D. dissertation, The University of California,<br />

Santa Cruz, CA. 114 pp.<br />

ROGOSA, D., AND G. GHANDOUR. 1991. Statistical models for behavioral observations.<br />

Journal of Educational Statistics 16: 157-252.<br />

SAMUELS, A., AND T. GIFFORD. 1997. A quantitative assessment of dominance relations<br />

among bottlenose dolphins. Marine Mammal Science 13:70-99.<br />

SCOTT, M. D., AND W. L. PERRYMAN.<br />

1991. Using aerial photogrammetry to study<br />

dolphin school structure. Pages 227-241 in K. Pryor and K. S. Norris, eds.<br />

Dolphin societies: Discoveries and puzzles. University of California Press, Berkeley,<br />

CA.<br />

SHANE, S. H. 1990. Behavior and ecology of the bottlenose dolphin at Sanibel Island,<br />

Florida. Pages 245-265 in S. Leatherwood and R. R. Reeves, eds. The bottlenose<br />

dolphin. Academic Press, Inc., New York, NY.<br />

TYACK, P. 1985. An optical telemetry device to identify which dolphin produces a<br />

sound. Journal of the Acoustical Society of America 78:1892-1895.<br />

TYACK, P. L., AND C. A. RECCHIA. 1991. A datalogger to identify vocalizing dolphins.<br />

Journal of the Acoustical Society of America 90: 1668-167 1.<br />

TYLER, S. 1979. Time-sampling: A matter of convention. Animal Behaviour 272301-<br />

810.<br />

WATERS, S., AND H. WHITEHEAD. 1990. Aerial behaviour in sperm whales. Canadian<br />

Journal of Zoology 68:2076-2082.<br />

WHITEHEAD, H. 1995. Investigating structure and temporal scale in social organizations<br />

using identified individuals. Behavioral Ecology 6: 199-208.<br />

WHITEHEAD, H. 1997. Analyzing animal social structure. Animal Behaviour 53: 1053-<br />

1067.<br />

WHITEHEAD, H., AND L. WEILGART. 1991. Patterns of visually observable behaviour<br />

and vocalizations in groups of female sperm whales. Behaviour 118:275-296.<br />

WILLIAMS, G. C. 1966. Adaptation and natural selection. Princeton University Press,<br />

Princeton, NJ.<br />

WURSIG, B., E. M. DORSEY, M. A. FRAKER, R. S. PAYNE AND W. J. RICHARDSON. 1985.<br />

Behavior of bowhead whales, Balaena mysticetlrs, summering in the Beaufort Sea:<br />

A description. Fishery Bulletin, U.S. 83:357-377.<br />

Received: 10 July 1995<br />

Accepted: 14 January 1998


MA”: BEHAVIORAL SAMPLING METHODS 121<br />

APPENDIX 1<br />

I conducted a search of the Science Citation Index (Institute for Scientific Infor-<br />

mation, Inc., 1996) using all 39 cetacean genera for keywords. The search yielded 846<br />

studies. To determine whether the study focused on cetacean behavior in the wild, each<br />

title was examined for keywords: behavior, acoustics, observations, field, captive, wild,<br />

and a description of the geographic location. The 846 studies were classified according<br />

to 14 subjects (e.g., molecular, neurophysiological, conservation, behavior); many stud-<br />

ies received more than one classification. Of the total, 125 studies (14.8%) in 28<br />

journals focused on cetacean behavior (n = 106) or vocalizations (n = 19) in the wild.<br />

Studies that appeared to focus strictly on population assessment (e.g., aerial counts)<br />

were not classified as behavioral studies. CJZ and MMS published 60% of the studies.<br />

The next most popular journal for cetacean behavior studies was the Journal ofMam-<br />

malogy, which published only 5% (n = 6). Thus, CJZ and MMS clearly represent the<br />

majority of relevant peer-reviewed publications in the Science Citation Index (SCI) and<br />

were chosen for in-depth analysis for this reason. Based on a comparison of Impact<br />

Factors (Impact Factor = the average number of times a paper in a journal is cited in<br />

the Science Citation Index within two years of publication), MMS and CJZ represent<br />

mid-level quality in research. Approximately an equal number of cetacean studies<br />

ranked above (n = 21) and below (n = 22) MMS and CJZ, with six studies (published<br />

in the Journal of Mammalogy) ranking the same as CJZ and one study ranking below<br />

CJZ but above MMS.<br />

Because the SCI search may have missed studies that did not use genera for key-<br />

words, I surveyed all issues of CJZ and MMS for all studies of free-ranging cetacean<br />

behavior and vocalizations published during 1989-1995. The SCI search by genera<br />

did not miss any studies, but one CJZ study I classified as a “behavior study” from<br />

the title was excluded from the review because the article’s content did not include<br />

behavior or vocalizations.<br />

The basic unit of analysis was one published article. Of 144 authors, 21 (14.6%)<br />

published more than one study. Of these, 14 published two studies, six published<br />

between three and five studies and one researcher authored or co-authored nine studies<br />

(using five different sampling methods).<br />

Species studied-Twenty-four cetacean species were studied in the 74 papers I re-<br />

viewed. Studies of humpback whales (Megaptera novaeangliae) were the most common<br />

(24%), followed by those of sperm whales (Physeter macrocephalus, 16%), and killer<br />

whales (Orcinus orca, 16%). Bottlenose dolphins (TurJiops truncatus) were represented in<br />

9% of the studies. Bowhead (Balaena mysticetus) and gray whales (Eschrictius robustus)<br />

were each represented in 5% of the studies. Narwhals (Monodon monoceros), Bryde’s<br />

whales (Balaenoptera edeni) and belugas (Delphinapterus leucas) were each represented in<br />

4% of the studies. Seven studies included more than one species. Of the 15 remaining<br />

species, only one or two studies were published on each.<br />

Age and sex classes sampled-Twenty studies focused on animals of a particular age<br />

or sex class. Otherwise, observers either did not know the ages or sexes of animals, or<br />

they observed animals of all ages and both sexes. No studies made quantitative com-<br />

parisons between age or sex classes.<br />

Number of animals sampled-The range of animals sampled was one to >1,000.<br />

Typically, researchers reported the number of animals that they surveyed or identified<br />

but not how many animals were observed or sampled. In 45% of the studies, the<br />

number of animals identified and/or sampled could not be determined. When groups<br />

of animals were sampled (e.g., in surveys or group-follows), 59% of the time (n=44<br />

studies) I could not distinguish between how many animals were individually identified<br />

and how many were actually observed.<br />

Individual identifiation-In 66% (n = 49) of the studies, researchers could identify<br />

individual animals. In 34% (n = 25), individual animals were not identified, although<br />

in about half of those (48%, n = 12), physical features of the species and reference to


122 MARINE MAMMAL SCIENCE, VOL. 15, NO. 1, 1999<br />

other studies indicated that individual identification was possible. In all studies that<br />

could identify individuals, rates of individual behavior or vocalizations were not re-<br />

ported unless the study involved only one or two animals (n = 3 studies).<br />

Observationirecording hours-Twenty-eight studies (38%) did not report how much<br />

time was spent observing or recording behavior. Some of these studies reported the<br />

hours spent at sea, but this information is only useful (as an indicator of sighting<br />

effort) when observation hours are also presented. Sixty-two percent of the studies<br />

reported the number of surveys conducted or the number of hours spent observing<br />

animals. Only four studies treated each subject, rather than each observation, as in-<br />

dependent for behavioral analysis.<br />

Group definitions and group size-Definition of “group” is essential to assess the va-<br />

lidity of behavioral sampling from groups. I indicated whether or not “group” was<br />

defined, and if it was defined, I noted the definition. Group-size definition was not<br />

relevant in nineteen studies (e.g., only one animal was studied). For 55 studies (surveys,<br />

group-follows, anecdotes), groups of animals were observed. In 42 of these (76%),<br />

researchers did not define “group.” If group was defined (n = 13), proximity-based<br />

measures (e.g., within 10 m) or coordinated behavior (all visible animals engaged in<br />

the same activity) were the two most common definitions of group.<br />

Group-size means or ranges (minimum and maximum group size) were reported in<br />

38 studies. Seventeen studies did not report group size information.<br />

Types of behavior and vocalizations recordd-Behaviors recorded in each study were<br />

classified as breathing (14%), diving (30%), resting (24%), surface displays (8%), so-<br />

cializing (30%), feeding or foraging (43%), traveling (27%), or “other.” Association<br />

(proportion of time animals together) was recorded in 22% of the studies and vocali-<br />

zations were recorded in 38%. “Other” included births (1%) or other unusual events<br />

(16%). Over half of the studies (57%) provided definitions for one or more of the<br />

behaviors recorded. Studies were classified according to the behaviors that researchers<br />

recorded, not which data were analyzed. If an ethogram was given, this was noted. No<br />

studies reported measures of inter-observer or intra-observer reliability.<br />

Sound recording was used either to find or track whales or to describe the vocal<br />

repertoire of a species, group, or individual. Of the 29 papers that recorded dolphin<br />

or whale vocalizations, 28% identified species-specific repertoires or group-specific rep-<br />

ertoires. In one study, species-typical vocalizations were monitored to test for the pres-<br />

ence of the species. Identification of individual repertoires within groups was limited<br />

to anecdotes, but one study used passive acoustic localization to identify which animal<br />

was likely to be vocalizing. Fourteen percent of the studies that recorded cetacean<br />

vocalizations calculated vocalization rates for groups of animals but not for individuals.<br />

Solitary individuals were acoustically recorded for portions of one study.<br />

All 74 studies were classified according to sampling method: ad libitum, continuous,<br />

focal-group, one-zero, point, scan, predominant activity, sequence, and incident sampling. If a<br />

study used more than one sampling method (six studies), both were scored. The results<br />

of these classifications are reported in the main text.


Aquatic Mammals 2008, 34(1), 35-43, DOI 10.1578/AM.34.1.2008.35<br />

Distribution and Habitat Use of Antillean Manatees<br />

(Trichechus manatus manatus) in the Drowned Cayes Area of<br />

Belize, Central America<br />

Katherine S. LaCommare, 1 Caryn Self-Sullivan, 2, 3 and Solange Brault 1<br />

1 University of Massachusetts, Boston, Department of Biology, 100 Morrissey Boulevard, Boston, MA 02125, USA;<br />

E-mail: kslacommare1@hotmail.com<br />

2 Texas A&M University, Department of Wildlife and Fisheries, College Station, TX 77845, USA<br />

3 <strong>Sirenian</strong> <strong>International</strong>, Inc., 200 Stonewall Drive, Fredericksburg, VA 22401, USA<br />

Abstract<br />

Belize, Central America, has long been recognized<br />

as a stronghold for Antillean manatees (Trichechus<br />

manatus manatus) in the Caribbean (O’Shea &<br />

Salisbury, 1991). The Drowned Cayes area, in<br />

particular, has been noted as an important habitat<br />

(Bengston & Magor, 1979; O’Shea & Salisbury,<br />

1991; Auil, 1998, 2004; Morales-Vela et al., 2000).<br />

It is critical to evaluate habitat use and the relative<br />

importance of different habitat types within these<br />

cayes because this area is increasingly impacted<br />

by human activities (Auil, 1998). The two research<br />

objectives for this paper are (1) to document<br />

manatee distribution within the Drowned Cayes,<br />

Swallow Caye, and Gallows Reef, and (2) to examine<br />

habitat use patterns in order to identify habitat<br />

characteristics influencing the probability of sighting<br />

a manatee. Binary logistic regression was used<br />

to examine whether the probability of sighting a<br />

manatee varied in relation to several habitat variables.<br />

The probability of sighting a manatee across<br />

all points was 0.31 per scan (n = 795). Habitat<br />

category, seagrass category, and habitat category<br />

interaction with resting hole were the most important<br />

variables explaining the probability of sighting<br />

a manatee. The Drowned Cayes area clearly constitutes<br />

a manatee habitat area. Seagrass flats and<br />

cove habitats with resting holes were especially<br />

important habitat characteristics.<br />

Key Words: distribution, habitat use, Antillean<br />

manatee, Trichechus manatus manatus, Belize<br />

Introduction<br />

The Antillean subspecies of the West Indian manatee<br />

(Trichechus manatus manatus) is found in<br />

19 countries throughout the Caribbean, Central<br />

America, and South America (Lefebvre et al.,<br />

2001). It is listed as vulnerable to extinction by<br />

the <strong>International</strong> Union for the Conservation of<br />

Nature and Natural Resources (IUCN) (2004)<br />

because of its greatly reduced numbers, continued<br />

exploitation, and population fragmentation.<br />

However, Belize, Central America, has long been<br />

recognized as a stronghold for Antillean manatees<br />

in the Caribbean (O’Shea & Salisbury, 1991). In<br />

the 1960s, Charnock-Wilson (1968, 1970) conducted<br />

interviews and personal observations and<br />

reported that manatees were abundant in the country.<br />

From the late 1970s to the early 2000s, aerial<br />

surveys continued to document that Belize has a<br />

relatively large number of manatees in comparison<br />

to neighboring countries (Bengston & Magor,<br />

1979; O’Shea & Salisbury, 1991; Auil, 1998,<br />

2004; Morales-Vela et al., 2000).<br />

The Drowned Cayes, Swallow Caye, and<br />

Gallows Reef, in particular, are recognized as<br />

important manatee areas within Belize. Charnock-<br />

Wilson (1968) described manatees using drowned<br />

cayes or mangrove islands, in a general sense,<br />

stating that it was common to spot manatees<br />

around drowned cayes, their channels, lagoons,<br />

and surrounding seagrass beds. Subsequent aerial<br />

surveys have documented high concentrations of<br />

manatees in the cayes near Belize City, of which<br />

the Drowned Cayes, Swallow Caye, and Gallows<br />

Reef are a part (Bengston & Magor, 1979; O’Shea<br />

& Salisbury, 1991; Auil, 1998, 2004; Morales-<br />

Vela et al., 2000). Boat surveys that have been<br />

conducted since 1999 further corroborated<br />

that this area is consistently used by manatees<br />

(LaCommare et al., 2003).<br />

This important manatee area may become<br />

increasingly threatened by tourism and human<br />

population growth (Auil, 1998); growth in tourism<br />

by cruise ships has been particularly dramatic.<br />

Between 1998 and 2006, the number of tourists<br />

entering Belize via cruise ships increased from<br />

14,183 visitors per year to 851,436 visitors per<br />

year (Belize Tourism Board, 2007). During that<br />

same period, there was an 18% increase in the<br />

population of Belize City (Brinkhoff, 2005). Due


36 LaCommare et al.<br />

to the Drowned Cayes’ proximity to Belize City,<br />

and because these islands lie between the mainland,<br />

the reef, and developed cayes, manatees could be<br />

particularly prone to threats from these sources. As<br />

tourism increases, a corresponding increase in boat<br />

traffic as well as land development is expected.<br />

Boats pass through the Drowned Cayes when traveling<br />

from Belize City to the northern cayes, outer<br />

cayes, and the reef, causing a risk of watercraft collisions<br />

with manatees. Watercraft collisions are the<br />

leading source of human-caused manatee mortality<br />

in Belize (Auil & Valentine, 2004). The proximity<br />

of these islands to the barrier reef may result in<br />

increased land development that could result in the<br />

loss of mangrove habitat. In the last five years, new<br />

resort development and expansion have occurred.<br />

And finally, development in Belize City and along<br />

the Belize River is likely to cause increases in runoff,<br />

which may decrease the productivity, biomass,<br />

and percent bottom cover of seagrass (Hemminga<br />

& Duarte, 2000; Duarte, 2002)—a principal food<br />

source for manatees (Ledder, 1986; Provancha &<br />

Hall, 1991; Mignucci-Giannoni, 1998; Lefebvre<br />

et al., 2000; U.S. Fish and Wildlife Service, 2001).<br />

Because the Drowned Cayes area is an important<br />

manatee area within the country and because the<br />

islands are increasingly affected by human activities,<br />

it is critical to improve our understanding of<br />

how manatees utilize this area. Evaluating animal<br />

habitat use and the relative importance of habitat<br />

types informs habitat management and conservation<br />

(Garshelis, 2000). The two research objectives<br />

for this paper are (1) to document manatee distribution<br />

within the Drowned Cayes, Swallow Caye,<br />

and Gallows Reef, and (2) to examine habitat use<br />

patterns in order to identify habitat characteristics<br />

influencing the probability of sighting a manatee.<br />

Materials and Methods<br />

Study Area<br />

The Drowned Cayes and Swallow Caye are mangrove<br />

islands along the central coast of Belize,<br />

10 to 15 km east of Belize City and 5 km west of<br />

the Belize Barrier Reef (Figure 1). The Drowned<br />

Cayes are a string of mostly uninhabited islands<br />

that are 14 km long by 4 km at their widest<br />

point and are almost entirely comprised of red<br />

(Rhizophora mangle) ) and black (Avicennia ( ger-<br />

minans) mangrove stands. There is very little dry<br />

land, and the islands are interspersed with broad<br />

channels, narrow inlets, shallow lagoons, and protected<br />

coves. The entire complex is surrounded by<br />

seagrass beds; seagrass also grows in varying densities<br />

on the bottoms of the channels, lagoons, and<br />

coves. Turtle grass (Thalassia testudinum) is the<br />

predominant species with shoal (Halodule wrightii)<br />

and manatee grass (Syringodium filiforme) also<br />

very common. Water depth is less than 1 m in<br />

some places and is never greater than 6 m within<br />

the study area; the tidal range is less than 0.3 m. In<br />

2002, the northern portion of the Drowned Cayes<br />

and Swallow Caye were designated as the Swallow<br />

Caye Wildlife Sanctuary.<br />

The islands are in a marine environment. In<br />

open water areas, salinities range from 35 to 40<br />

ppt (mean salinity is 37 ppt), making the environment<br />

slightly hypersaline. Within the mangrove<br />

island complex, salinities have greater fluctuations<br />

as a result of larger run-off, evaporation, and tidal<br />

influences. Therefore, within the channels, coves,<br />

and lagoons of the mangrove islands, salinities<br />

can range from 30 to 42 ppt.<br />

Survey Design<br />

A point sampling survey design was devised for a<br />

small boat platform to quantify manatee distribution<br />

and habitat use. Fifty-four permanent points<br />

were established in all habitat types throughout<br />

the study area (Figure 2). Habitat types are<br />

defined in Table 1. These points were randomly<br />

sampled during January, February, March, June,<br />

July, and August from 2001 to 2004. Surveys were<br />

also conducted in June, July, and August 2005.<br />

Using a mix of experienced observers and volunteers,<br />

three to 13 observers searched for manatees<br />

at each point for 30 min. These searches are<br />

referred to as point scans. As the boat came within<br />

100 m of each point, observers started scanning<br />

for manatees in a 360º circle around the boat. The<br />

boat was then anchored in position using a pole.<br />

For each scan, the number of manatees and habitat<br />

characteristics were recorded. The latter consisted<br />

of habitat category, presence of a resting<br />

hole, presence and type of seagrass, temperature,<br />

salinity, and sea state (Table 2).<br />

Data Analysis<br />

Manatee Distribution—Manatee distribution was<br />

mapped within the Drowned Cayes area by calculating<br />

the probability of sighting a manatee for each<br />

point. Since most of the scans had no manatees and<br />

because location sample sizes were unequal, presence<br />

or absence of manatees was used to calculate<br />

the probability of sighting a manatee. Points that<br />

were visited less than five times were excluded<br />

from this analysis. This map is a snapshot of the<br />

overall variation in distribution throughout the<br />

study area. To provide the most comprehensive<br />

map of distribution, the largest number of points<br />

possible was included in the analysis. To do this,<br />

20-min scans were included in the analysis. To<br />

create parity between the 20- and 30-min scans,<br />

presence or absence of the 30-min scans was<br />

determined based on only the first 20 min of the<br />

sampling duration. This increased the sample size


Distribution and Habitat Use of Antillean Manatees in Belize 37<br />

Figure 1. Map of Drowned Cayes and surrounding area (Source: British Admiralty Navigation Chart No. 522, Belize City<br />

and Approaches, 1989)<br />

from 613 to 795 scans and increased the number of<br />

points used in the analysis from 39 to 47.<br />

Habitat Use—Binary logistic regression was used<br />

to examine whether the probability of sighting a manatee<br />

varied in relation to several habitat variables—<br />

habitat category, presence of a resting hole, presence<br />

and type of seagrass, temperature, salinity, and sea<br />

state. Due to the large number of zeros (268 zero<br />

counts of 430 samples) and unequal sample sizes<br />

across habitat categories, manatee presence/absence<br />

was used rather than the number of individual manatees<br />

sighted during each scan to calculate the probability<br />

of sighting a manatee. Habitat category was<br />

determined by placing each point into one of six<br />

categories based on mangrove shoreline features and<br />

the depth profile of the particular point. These habitat<br />

categories were qualitatively assigned and are mutually<br />

exclusive. It is recognized that not all points fit<br />

neatly into these designations. A complete description<br />

of each category is given in Table 1. A resting<br />

hole is a bottom feature that is a distinct, shallow<br />

depression in the seafloor. Although they are at least<br />

3 to 4 m wide by 3 to 4 m long, they can be larger and<br />

are not necessarily regularly shaped. In some cases,<br />

they appear to be natural features that are maintained<br />

by manatee use; in other cases, they may have been<br />

created by manatee use. Seagrass category connotes<br />

whether seagrass was present at a particular point<br />

and, if so, what species of seagrass were found there.<br />

There were six categories: (1) none, (2) Thalassia


38 LaCommare et al.<br />

Figure 2. Map of manatee sighting probability throughout the Drowned Cayes study area (n = 795) (Source: Defense<br />

Mapping Agency Chart, Belize City Harbor, 1996)


Distribution and Habitat Use of Antillean Manatees in Belize 39<br />

Table 1. List and description of habitat categories used in the logistic regression analysis; number of points equals the number<br />

of points in each habitat category type. Description is the definition of the habitat category type.<br />

Habitat category<br />

Table 2. Description and assessment of the model relating the presence/absence of manatees to habitat variables for the<br />

Drowned Cayes study area (n = 491, Nagelkerke R-Square = 0.22); significant variables were habitat category, seagrass<br />

category, and habitat category*resting hole interaction term. Backward stepwise procedure and log-likelihood function were<br />

used to determine the most parsimonious model. Hosmer and Lemeshow Goodness of Fit χ 2 = 3.757, p = .878.<br />

Variables in the model<br />

testudinum only, (3) Halodule wrightii only,<br />

(4) T. testudinum/H. /H. / wrightii mix, (5) Syringodium<br />

filiforme/T. testudinum mix, and (6) a mix of all<br />

three. Sea state was based on the Beaufort scale.<br />

Logistic regression accommodates both continuous<br />

and categorical predictor variables (Trexler<br />

& Travis, 1993; Floyd, 2001). Maximum likelihood<br />

method was used to fit the model to the data,<br />

and a backward stepwise procedure was used to<br />

determine the most parsimonious model. The likelihood<br />

ratio test and the Wald statistic were used<br />

to determine the significance of the parameters in<br />

explaining the variation in the dependent variable.<br />

To determine if the model was an adequate fit to the<br />

Change in -2 log<br />

likelihood df<br />

data and to rule out the undue influence of outliers,<br />

the Hosmer-Lemeshow test was used and observed<br />

and expected sighting probabilities were compared<br />

(Trexler & Travis, 1993). Plots of Cook’s distances<br />

and Studentized residuals vs predicted probabilities<br />

were used to examine the influence of outliers and<br />

bias in the data. All statistical analyses were performed<br />

using SPSS, Version 13.0 (SPSS, 2004).<br />

Results<br />

Significance<br />

of change<br />

Dependent variable<br />

Manatee presence/absence<br />

Independent variables<br />

Sea state (Beaufort scale) 0.55 3 NS<br />

Surface salinity (ppt) 1.76 1 NS<br />

Water temperature (ºC) 0.11 1 NS<br />

Habitat category 14.09 5 0.015*<br />

Resting hole 1.27 1 NS<br />

Seagrass category 23.21 5 0.0001*<br />

Habitat category*resting hole 10.46 3 0.015*<br />

Habitat category*seagrass category 10.84 8 NS<br />

NS = Not significant, * = significant<br />

Number<br />

of points Description<br />

Lagoon 5 A large open water area totally encompassed within the mangrove islands; the area<br />

has a uniform and shallow depth (< 3 m).<br />

Channel 12 An area of deeper water (3 to 6 m) that cuts between the mangrove islands;<br />

generally, there is mangrove shoreline on two sides.<br />

Channel edge 8 An area of deeper water that cuts through areas of shallower water; the point<br />

encompasses two habitat types: (1) channel and (2) seagrass bed. Depth ranges from<br />

1 to 5 m. The point may be adjacent to mangrove shoreline on one side or may be in<br />

open water.<br />

Seagrass bed 11 An area of shallow water, < 3 m, outside of the mangrove islands with a seagrass<br />

bottom.<br />

Cove 16 An area of very protected water that is at the end of a channel or off to the side of a<br />

channel; it is nearly enclosed by mangroves and has shoreline on at least three sides.<br />

Depth ranges from 0.6 to 4 m.<br />

Reef 2 This is an open water area on the back reef portion of the Belize Barrier Reef; there<br />

are no mangrove islands in the vicinity, and depth ranges from 1.5 to 4 m.<br />

Distribution Within the Drowned Cayes<br />

Manatees were sighted throughout the study area.<br />

The probability of sighting a manatee across all


40 LaCommare et al.<br />

points was 0.31/20-min scan (n = 795). Sighting<br />

probabilities for each location ranged from 0 to<br />

0.86/scan. Swallow Caye had the highest probability<br />

of sightings among its points, and the<br />

Gallows Reef points had the lowest probability of<br />

sightings. There does not appear to be a distinct<br />

spatial pattern of variability in sighting probability<br />

among the Drowned Cayes points (Figure 2).<br />

Habitat Use<br />

The probability of sighting a manatee was lowest<br />

at the reef (0.<strong>17</strong>/scan) and highest on seagrass<br />

flats (0.58/scan) (Figure 3). Habitat category,<br />

seagrass category, and habitat category interaction<br />

with resting hole were the most important<br />

variables explaining the probability of sighting<br />

a manatee (Table 2). The change in log likelihood<br />

for each of these three terms was 14.09 with<br />

p = 0.015, 23.21 with p = 0.0001, and 10.46 with<br />

p = 0.015, respectively. Within habitat categories,<br />

seagrass flats and coves with resting holes both<br />

explained a significant portion of the variation in<br />

manatee presence (Wald = 4.14 with p = 0.042<br />

and 4.79 with p = 0.029, respectively) (Figures<br />

3A & B). There was only one scan point categorized<br />

as a channel with a resting hole. The low<br />

sample size of this category type may explain<br />

why channels with resting holes had an opposite<br />

trend to other habitat types. Scan points with<br />

just T. testudinum and S. filiforme present or with<br />

just H. wrightii present had a significantly lower<br />

probability of sighting a manatee than other seagrass<br />

categories (Wald = 6.50 with p = 0.011 and<br />

6.4 with p = 0.011, respectively) (Figure 4). The<br />

Hosmer-Lemeshow chi-square goodness of fit<br />

was 1.486 with a p-value of 0.983, indicating that<br />

the overall model was a good fit to the data, and<br />

the Nagelkerke R-Square was 0.221 (Table 2).<br />

The expected sighting probability for each habitat<br />

and seagrass category was very similar to the<br />

observed sighting probability further indicating<br />

that the model was a good fit to the data (Figures<br />

3A & 4). Residual plots of change in deviance vs<br />

predicted probabilities indicated that samples with<br />

a low predicted probability of a presence were<br />

poorly fit by the model. In other words, manatees<br />

tended to be observed more often than predicted<br />

by the model at places where the probability of<br />

sighting a manatee was low.<br />

Discussion<br />

The Drowned Cayes area clearly constitutes a<br />

manatee habitat area. Manatees were sighted at<br />

least once at nearly all of the points that were<br />

surveyed (47 out of 54 points). There were<br />

some points that had particularly high sighting<br />

probabilities (≥ 0.5/20-min scan), and these points<br />

A.<br />

B.<br />

Figure 3. (A) Observed and predicted probability of sighting<br />

a manatee per 20-min scan by habitat; (B) Observed<br />

probability of sighting a manatee per 20-min scan by habitat<br />

with and without the presence of a resting hole; No =<br />

no resting hole, Yes = resting hole present (2001 to 2005,<br />

n = 491, ± 1 SE).<br />

Figure 4. Observed and predicted probability of sighting<br />

a manatee per 20-min scan by seagrass category (2001 to<br />

2005, n = 491)<br />

might be considered “hot spots” within the overall<br />

mangrove island complex. These points should<br />

be given particular consideration in management<br />

plans in terms of tourist activities, development,<br />

and watercraft traffic.


Distribution and Habitat Use of Antillean Manatees in Belize 41<br />

Based on the logistic regression procedure,<br />

habitat category was one of the three most important<br />

variables explaining variation in sighting<br />

probability across the Drowned Cayes. Variability<br />

in sighting conditions between habitat types did<br />

not result in differences in detection probability.<br />

It is unlikely that there were detection biases as a<br />

result of habitat types (LaCommare et al., unpub.<br />

data). Seagrass beds had the highest probability<br />

of sighting a manatee, and reefs had the lowest.<br />

Seagrass flats may be a particularly important<br />

habitat category, indicating the importance of the<br />

Drowned Cayes as a feeding area.<br />

However, manatees did utilize all habitat categories.<br />

Manatees use the entire area and many of<br />

its components to meet a variety, but probably not<br />

all, of their physiological and behavioral requirements.<br />

During the course of this study, manatees<br />

were observed feeding, socializing, resting,<br />

nursing calves, and moving from place to place<br />

(Self-Sullivan & LaCommare, unpub. data). Even<br />

resources that are used at a low frequency may be<br />

important components of the manatees’ overall<br />

habitat. For instance, the reef may be a seasonally<br />

important travel corridor and stopover site for male<br />

manatees during the mating season (Self-Sullivan<br />

et al., 2004). Behavioral and physiological studies<br />

indicate that manatees probably need regular<br />

access to freshwater sources to survive (Reynolds<br />

& Odell, 1991; Ortiz et al., 1998, 1999; Reynolds,<br />

1999; Deutsch et al., 2000, 2003). Only one source<br />

of hyposaline water (refractometer measurements<br />

ranged from 10 to <strong>17</strong> ppt) has been conclusively<br />

identified (Self-Sullivan & LaCommare, unpub.<br />

data) within the Drowned Cayes, which may be<br />

used for osmoregulation. Manatees may also be<br />

traveling from the Drowned Cayes to freshwater<br />

sources several kilometers away as they do in<br />

Florida (Deutsch et al., 2003). Management plans<br />

should recognize the importance of protecting a<br />

variety of habitat types.<br />

Seagrass is clearly an important component of<br />

the Drowned Cayes habitat. Seagrass category<br />

was an important variable explaining the probability<br />

of sighting a manatee. It contributed the<br />

largest improvement in model fit, and the parameter<br />

was highly significant. Specifically, sites that<br />

had a mix of T. testudinum and S. filiforme or sites<br />

with just H. wrightii present had a significantly<br />

lower probability of sighting a manatee than the<br />

other seagrass categories—no seagrass, a mix of<br />

all three species, just T. testudinum, , or a mix of T.<br />

testudinum and H. wrightii.<br />

How this relates to manatee foraging and feeding<br />

is not clear. The three most common species<br />

in the study area—T. testudinum, H. wrightii,<br />

and S. filiforme—are present in the manatee<br />

diet in both Florida and Puerto Rico (Packard,<br />

1984; Ledder, 1986; Provancha & Hall, 1991;<br />

Mignucci-Giannoni, 1998; Lefebvre et al., 2000;<br />

USFWS, 2001). The relative importance of these<br />

species in their diet is not fully understood in<br />

these places. In Indian River Lagoon, Florida, and<br />

in Puerto Rico, “manatees fed most often on the<br />

most frequently encountered seagrass” (Lefebvre<br />

et al., 2000, p. 295). In Indian River Lagoon, this<br />

was H. wrightii, and in Puerto Rico, this was T.<br />

testudinum (Lefebvre et al., 2000). However, in<br />

both locations, it is possible that manatees return<br />

to specific H. wrightii beds to feed on them<br />

(Lefebvre et al., 2000). Both T. testudinum and<br />

H. wrightii appear to be important food resources<br />

in the Drowned Cayes based on observations of<br />

feeding manatees, and certain areas appear to be<br />

more heavily foraged than others (LaCommare<br />

& Self-Sullivan, unpub. data), perhaps indicating<br />

foraging preferences.<br />

While certain types of seagrass sites, as well as<br />

species, may be particularly important to manatees<br />

in the Drowned Cayes, fully understanding<br />

manatee foraging will involve examining the<br />

extent of feeding behavior in particular places,<br />

specific physical and biological characteristics of<br />

important seagrass areas, and the juxtaposition of<br />

these areas to other important resources.<br />

Habitat category interacting with resting hole<br />

was also an important variable explaining variation<br />

in sighting probability. Cove habitats with<br />

resting holes were the most significant subcategory.<br />

Cove habitats are extremely sheltered places<br />

found at the end of narrow channels or off to the<br />

side of larger channels. In the Drowned Cayes,<br />

resting holes are a feature of the bottom topography<br />

and are associated with resting manatees<br />

(Self-Sullivan & LaCommare, unpub. data). In<br />

Florida, manatees often use secluded places for<br />

feeding, resting, mating, and calving (USFWS,<br />

2001). This appears to be true in the Drowned<br />

Cayes as well. The presence of a resting hole was<br />

a key feature distinguishing which particular cove<br />

area was utilized (0.25 to 0.49/scan increase in<br />

sighting probability). It is unclear at this point if<br />

the resting hole was the reason manatees used the<br />

site or if the resting hole depression is a result of<br />

repeated manatee use.<br />

While the Nagelkerke R-Square for the regression<br />

model was 0.22, indicating that only 22% of<br />

the variability in sighting probability was explained<br />

by the logistic regression model, the model was a<br />

good fit to the data, and the variability explained<br />

was the correct variability. The ability to explain<br />

variation in sighting probability will provide a<br />

better understanding of habitat use. Building a<br />

habitat model that includes variables that describe<br />

the spatial configuration and characteristics of habitat<br />

is an important follow-up to this analysis and


42 LaCommare et al.<br />

may even provide insight into habitat selection and<br />

preferences. Habitat use is dependent on behavior,<br />

and suitable habitat must include a variety of patch<br />

types or components so that animals can meet their<br />

essential behavioral and physiological requirements<br />

(Mysterud & Ims, 1998). In addition, habitat<br />

use, selection, and preferences are affected by habitat<br />

heterogeneity (Johnson, 1980; Li & Reynolds,<br />

1994; Kie et al., 2002). Heterogeneity can be<br />

defined in terms of number, proportion, spatial<br />

arrangement, shape, and contrast between patches<br />

and patch types (Kie et al., 2002). Analyses examining<br />

manatee habitat components in these terms<br />

and with respect to behavior could be important in<br />

understanding manatee habitat use patterns.<br />

Acknowledgments<br />

We thank Earthwatch Institute for substantial<br />

financial support and Earthwatch Volunteers for<br />

logistical support and data collection. We thank<br />

Conservation Action Fund, Henry Greenwalt,<br />

Project Aware, Virtual Explorers, and Jack Bert<br />

for additional financial support. Spanish Bay<br />

Resort provided logistical and in kind support<br />

from 2001 to 2004. Hugh Parkey Foundation for<br />

Marine Awareness and Education provided logistical<br />

and in kind support in 2005. We are grateful<br />

to the numerous interns and field assistants who<br />

were vital to data collection and field logistics. We<br />

are especially thankful for the help and guidance<br />

of our colleague, friend, and field assistant Gilroy<br />

Robinson. In addition, we thank the two anonymous<br />

reviewers for their insightful comments that<br />

led to the improvement of this manuscript.<br />

Literature Cited<br />

Auil, N. E. (1998). Belize manatee recovery plan (United<br />

Nations Development Program/Global Environment<br />

Facility, Coastal Zone Management Project No. BZE/92/<br />

G31). Kingston, Jamaica: United Nations Environment<br />

Program, Caribbean Environment Program.<br />

Auil, N. E. (2004). Abundance and distribution trends of<br />

the West Indian manatee in the coastal zone of Belize:<br />

Implications for conservation. Unpublished Master’s<br />

thesis, Texas A&M University, College Station. 83 pp.<br />

Auil, N. E., & Valentine, A. (2004). Manatee strandings<br />

along the coastal zone of Belize 1996-2003. Unpublished<br />

report, Coastal Zone Management Authority and<br />

Institute.<br />

Belize Tourism Board. (2007). Belize tourism statistics.<br />

Retrieved 11 March 2008 from www.belizetourism.org/<br />

belize-tourism/tourist-arrivals-by-mode-of-arrival.html.<br />

Bengston, J. L., & Magor, D. (1979). A survey of manatees<br />

in Belize. Journal of Mammalogy, 60(1), 230-232.<br />

Brinkhoff, T. (2005). City population. Retrieved 29<br />

February 2008 from www.citypopulation.de.<br />

Charnock-Wilson, J. (1968). The manatee in British<br />

Honduras. Oryx, 8(4), 293-294.<br />

Charnock-Wilson, J. (1970). Manatees and crocodiles.<br />

Oryx, 10(4), 236-238.<br />

Deutsch, C. J., Reid, J. P., Bonde, R. K., Easton, D. E.,<br />

Kochman, H. I., & O’Shea, T. J. (2000). Seasonal<br />

movements, migratory behavior, and site fidelity of<br />

West Indian manatees along the Atlantic coast of<br />

the United States as determined by radio-telemetry.<br />

Unpublished report, Florida Cooperative Fish and<br />

Wildlife Research Unit, USGS and University of<br />

Florida.<br />

Deutsch, C. J., Reid, J. P., Bonde, R. K., Easton, D. E.,<br />

Kochman, H. I., & O’Shea, T. J. (2003). Seasonal<br />

movements, migratory behavior, and site fidelity of<br />

West Indian manatees along the Atlantic coast of the<br />

United States. Wildlife Monographs, 67, 1-77.<br />

Duarte, C. M. (2002). The future of seagrass meadows.<br />

Environmental Conservation, 29(2), 192-206.<br />

Floyd, T. (2001). Logit modeling and logistic regression. In<br />

S. M. Scheiner & J. Gurevitch (Eds.), Design and analysis<br />

of ecological experiments (pp. 197-216). New York:<br />

Oxford University Press. 415 pp.<br />

Garshelis, D. L. (2000). Delusions in habitat evaluation:<br />

Measuring use, selection and importance. In L. Boitani<br />

& T. K. Fuller (Eds.), Research techniques in animal<br />

ecology (pp. 111-164). New York: Columbia University<br />

Press. 442 pp.<br />

Hemminga, M. A., & Duarte, C. M. (2000). Seagrass ecology.<br />

Cambridge, UK: Cambridge University Press.<br />

<strong>International</strong> Union for the Conservation of Nature and<br />

Natural Resources (IUCN). (2004). 2004 IUCN red list<br />

of threatened species. Retrieved 16 November 2005<br />

from www.redlist.org.<br />

Johnson, D. (1980). The comparison of usage and availability<br />

measurements for evaluating resource preference.<br />

Ecology, 61(1), 65-71.<br />

Kie, J. G., Bowyer, R. T., Nicholson, M. C., Boroski, B. B.,<br />

& Loft, E. R. (2002). Landscape heterogeneity at differing<br />

scales: Effects on spatial distribution of mule deer.<br />

Ecology, 83(2), 530-544.<br />

LaCommare, K. S., Self-Sullivan, C., & Brault, S. (2003).<br />

Habitat use of Antillean manatees (Trichechus manatus<br />

manatus) in the Drowned Cayes, Belize, Central<br />

America. Paper presented at the Fifteenth Biennial<br />

Conference on the Biology of Marine Mammals.<br />

Vancouver, Washington.<br />

Ledder, D. A. (1986). Food habits of the West Indian manatee,<br />

Trichechus manatus latirostris, in south Florida.<br />

Coral Gables, FL: University of Miami.<br />

Lefebvre, L. W., Reid, J. P., Kenworthy, W. J., & Powell,<br />

J. A. (2000). Characterizing manatee habitat use and seagrass<br />

grazing in Florida and Puerto Rico: Implications<br />

for conservation and management. Pacific Conservation<br />

Biology, 5, 289-298.<br />

Lefebvre, L. W., Marmontel, M., Reid, J. P., Rathbun,<br />

G. B., & Domning, D. P. (2001). Status and biogeography<br />

of the West Indian manatee. In C. A. Woods &


Distribution and Habitat Use of Antillean Manatees in Belize 43<br />

F. E. Sergile (Eds.), Biogeography of the West Indies:<br />

Patterns and perspectives (pp. 425-474). Boca Raton,<br />

FL: CRC Press.<br />

Li, H., & Reynolds, J. F. (1994). A simulation experiment<br />

to quantify spatial heterogeneity in categorical maps.<br />

Ecology, 75(8), 2446-2455.<br />

Mignucci-Giannoni, A. A. (1998). The diet of the manatee<br />

(Trichechus manatus) in Puerto Rico. Marine Mammal<br />

Science, 14(2), 394-397.<br />

Morales-Vela, B., Olivera-Gomez, D., Reynolds, J. E., III,<br />

& Rathbun, G. B. (2000). Distribution and habitat use<br />

by manatees (Trichechus manatus manatus) in Belize<br />

and Chetumal Bay, Mexico. Biological Conservation,<br />

95, 67-75.<br />

Mysterud, A., & Ims, R. A. (1998). Functional responses in<br />

habitat use: Availability influences relative use in tradeoff<br />

situations. Ecology, 79(4), 1435-1441.<br />

Ortiz, R. M., Worthy, G. A. J., & Byers, F. M. (1999).<br />

Estimation of water turnover rates of captive West<br />

Indian manatees (Trichechus manatus) held in fresh and<br />

salt water. The Journal of Experimental Biology, 202,<br />

33-38.<br />

Ortiz, R. M., Worthy, G. A. J., & MacKenzie, D. S. (1998).<br />

Osmoregulation in wild and captive West Indian manatees<br />

(Trichechus manatus). Physiological Zoology,<br />

71(4), 449-457.<br />

O’Shea, T. J., & Salisbury, C. A. L. (1991). Belize: A last<br />

stronghold for manatees in the Caribbean. Oryx, 25(3),<br />

156-164.<br />

Packard, J. M. (1984). Impact of manatees Trichechus<br />

manatus on seagrass communities in eastern Florida.<br />

Acta Zoologica Fennica, <strong>17</strong>2, 21-22.<br />

Provancha, J. A., & Hall, C. R. (1991). Observations of<br />

associations between seagrass beds and manatees in<br />

East Central Florida. Florida Scientist, 54(2), 86-98.<br />

Reynolds, J. E., III. (1999). Efforts to conserve manatees.<br />

In J. R. Twiss, Jr. & R. R. Reeves (Eds.), Conservation<br />

and management of marine mammals (pp. 267-293).<br />

Washington, DC: Smithsonian Institution Press.<br />

Reynolds, J. E., III, & Odell, D. K. (1991). Manatees and<br />

dugongs. New York: Facts on File, Inc.<br />

Self-Sullivan, C., Smith, G. W., Packard, J. M., &<br />

LaCommare, K. S. (2004). Seasonal occurrence of<br />

male Antillean manatees (Trichechus manatus manatus)<br />

on the Belize Barrier Reef. Aquatic Mammals, 29(3),<br />

342-354.<br />

SPSS. (2004). SPSS release 13.0 [Computer software].<br />

Apache Software Foundation.<br />

Trexler, J. C., & Travis, J. (1993). Nontraditional regression<br />

analyses. Ecology, 74(6), 1629-1637.<br />

U.S. Fish and Wildlife Service (USFWS). (2001).<br />

Technical/Agency draft: Florida manatee recovery plan<br />

(Trichechus manatus latirostris). Atlanta, GA: Author.


<strong>17</strong>2<br />

NOTES<br />

Caribbean Journal of Science, Vol. 41, No. 1, <strong>17</strong>2-<strong>17</strong>7, 2005<br />

Copyright 2005 College of Arts and Sciences<br />

University of Puerto Rico, Mayagüez<br />

Bottlenose Dolphins (Tursiops<br />

truncatus) in the Drowned Cayes,<br />

Belize: Group Size, Site Fidelity<br />

and Abundance<br />

KECIA A. KERR 1 ,R.H.DEFRAN 2 , AND GRE-<br />

GORY S. CAMPBELL Cetacean Behavior Laboratory<br />

- Department of Psychology, San Diego<br />

State University, San Diego, CA 92182-4611,<br />

USA<br />

ABSTRACT.—Group size, site fidelity and abundance<br />

of bottlenose dolphins, Tursiops truncatus,<br />

ms. submitted June 3, 2004; accepted December 28,<br />

2004<br />

1<br />

Current address: Whale Research Laboratory, Department<br />

of Geography, University of Victoria, PO<br />

Box 3050 STN CSC, Victoria, BC, V8W 3P5 Canada<br />

2<br />

Corresponding author: rdefran@sunstroke.<br />

sdsu.edu


were assessed during 392 photo-identification surveys<br />

conducted during 1997-1999 in the Drowned<br />

Cayes region, near Belize City, Belize, Central<br />

America. During this study 2155 dolphins were<br />

sighted across 736 groups. Mean group size was 2.9<br />

(SD = 2.32) which is one of the smallest reported for<br />

bottlenose dolphins. One hundred and fifteen individual<br />

dolphin were photographically identified,<br />

with sighting frequencies ranging from one to fifty<br />

(X ¯ = 8.1, SD = 9.05). Thirty percent of identified dolphins<br />

were judged to be residents, while 23% were<br />

photographed only once. Chao’s M th model for<br />

closed populations was used to derive an abundance<br />

estimate of 122 dolphins (95% CI = 114 -140). This<br />

low abundance estimate and a leveling trend in the<br />

rate of newly identified individuals, indicates that<br />

the Drowned Cayes dolphin population is both<br />

small and finite. Group size, abundance, and site<br />

fidelity comparisons were made with a 4-yr photoidentification<br />

study conducted at nearby Turneffe<br />

Atoll. Both the Drowned Cayes and Turneffe Atoll<br />

studies had similarly small group sizes, low and<br />

variable levels of site fidelity and low abundance<br />

estimates, but there was no overlap between individual<br />

sightings in the two areas. The observed behavioral<br />

patterns and similarities between the two<br />

studies raise concerns that increasing pressures on<br />

Belize’s marine resources may pose a threat to its<br />

bottlenose dolphins.<br />

KEYWORDS.—Residency patterns, population dynamics,<br />

cetaceans, Belize, Caribbean<br />

A recent 4-yr study at Turneffe Atoll in<br />

Belize, Central America (Campbell et al.<br />

2002), was one of the first to document the<br />

population dynamics of bottlenose dolphins<br />

(Tursiops truncatus) in a tropical offshore<br />

atoll containing coral reefs, sea grass<br />

beds and mangroves, and is one of the few<br />

published studies on this cetacean species<br />

in the Caribbean Sea (see also Grigg and<br />

Markowitz 1997; Rossbach and Herzing<br />

1999; Rogers et al. 2004). Turneffe bottlenose<br />

dolphins occurred in small groups<br />

that were larger when they contained<br />

calves, and this population had a high proportion<br />

of individuals sighted only once.<br />

Further, abundance estimates were low,<br />

and only a small proportion of these dolphins<br />

showed evidence of site fidelity<br />

(Campbell et al. 2002). The habitat characteristics<br />

of the Drowned Cayes, where the<br />

current study was conducted, are similar to<br />

those of Turneffe. Thus, the objectives of<br />

NOTES <strong>17</strong>3<br />

the current study were to examine the generality<br />

of the Turneffe findings and interpretations,<br />

identify possible overlap between<br />

these two dolphin populations, and<br />

extend the management and conservation<br />

implications of both studies.<br />

The Drowned Cayes are located in the<br />

western Caribbean, south of the Yucatan<br />

Peninsula, 6 km east of Belize City and 16<br />

km west of Turneffe Atoll (Fig. 1). The<br />

study area consisted of a 150 km 2 chain of<br />

mangrove islands extending from <strong>17</strong>°20’N-<br />

<strong>17</strong>°34’N, and 88°03’W-88°07’W. Like<br />

Turneffe Atoll, the Drowned Cayes area is<br />

characterized by a predominating substrate<br />

of sandy sea grass beds, but also includes a<br />

portion of the Belize Barrier Reef. Surface<br />

water temperature ranged from 25°-33°C<br />

(X ¯ = 29.2, SD = 1.62), and water depth, measured<br />

during 216 sightings, ranged from 0.9<br />

mto11.8m(X ¯ = 4.3 m). All aspects of the<br />

photo-identification methodology used in<br />

this study, including: survey methodology,<br />

criteria for groups and calves, image processing<br />

and analysis of photographic data<br />

(i.e., rate of discovery, sighting frequency,<br />

abundance), were the same as those described<br />

in Campbell et al. (2002). Briefly<br />

summarized, boat based surveys lasting<br />

approximately 4 h were carried out once or<br />

twice daily, and all portions of the study<br />

area were covered at least once a month.<br />

When dolphins were sighted, the survey<br />

vessel stopped while environmental data<br />

were recorded and group size was determined.<br />

Next, an attempt was made to obtain<br />

high quality dorsal fin photographs for<br />

each group member. Once photographic<br />

data were collected, the survey continued<br />

along a predetermined route to search for<br />

and photograph additional dolphins.<br />

During 1246 h expended over a 3-yr period<br />

(February-December in 1997 and 1998,<br />

and April-Dec in 1999), 392 surveys were<br />

completed, and 277 h were spent in direct<br />

observation of 2155 dolphins sighted across<br />

736 groups. Group size ranged from one to<br />

20 dolphins (X ¯ = 2.9, SD = 2.32), but was<br />

higher for groups with calves (n = 160, X ¯ =<br />

4.6 [includes calves], SD = 2.68) than without<br />

calves (n = 576, X ¯ = 2.5, SD =<br />

1.99)(Mann-Whitney U = 19540.0, Z =<br />

-11.433, p = < 0.001). Almost a third (31%)


<strong>17</strong>4<br />

FIG. 1. Drowned Cayes study area (within dashed outline). Insets show the location of the study area from a<br />

broad regional (top) and western Caribbean (bottom) perspective.<br />

of all sightings were solitary dolphins and<br />

99% of all groups contained 10 or fewer<br />

dolphins. Group sizes for Drowned Cayes<br />

(X ¯ = 2.9) and Turneffe dolphins (X ¯ = 3.8, SD<br />

= 3.55, Campbell et al. 2002) are among the<br />

smallest reported for any population of<br />

bottlenose dolphins (Connor et al. 2000).<br />

Group size in bottlenose dolphins, as in<br />

many other species, is a tradeoff between<br />

optimizing foraging efficiency and minimizing<br />

predation risk (Wells and Scott<br />

1999; Campbell et al. 2002). While there<br />

have been no studies of prey characteristics<br />

for Drowned Cayes or Turneffe dolphins to<br />

date, no observations or images indicated<br />

evidence of shark bites, suggesting that<br />

predatory threats from sharks are minimal.<br />

Low apparent predation at the Drowned<br />

Cayes suggests, as it did at Turneffe, that<br />

energy intake is a primary selective pressure<br />

on group size; and, the small group<br />

sizes in these areas are optimized for food<br />

resources that are likely low in density.<br />

NOTES<br />

Similarly, in both the Drowned Cayes and<br />

Turneffe, larger calf-groups may be an adaptation<br />

that facilitates allomaternal behavior,<br />

thereby increasing the foraging efficiency<br />

of nursing mothers who are<br />

constrained by maternal responsibilities<br />

(see review in Campbell et al. 2002).<br />

Across the study period, 115 dolphins<br />

were photographically identified. The<br />

slope of the Drowned Cayes discovery<br />

curve for newly photographed dolphins<br />

showed a steep rise until the 90 th survey<br />

when 76 dolphins (66%) had been identified<br />

(compare with Fig. 2, Campbell et al.<br />

2002). Few new dolphins were photographed<br />

until surveys 190-230 (Aug 1999)<br />

when an additional 23 dolphins (18%) were<br />

identified. No new dolphins were photographed<br />

during the remaining 162 surveys.<br />

Sighting frequencies for identified dolphins<br />

ranged from 1-50 (X ¯ = 8.1, SD = 9.02), with<br />

23% (n = 27) of these dolphins identified<br />

one time, 50% (n = 57) sighted between two


and nine times and 27% (n = 31) sighted 10<br />

or more times. Evidence for site fidelity<br />

was evaluated by examining sighting frequencies<br />

within and between the three<br />

study years. Dolphins sighted two or more<br />

times in each of the three study years, or<br />

four or more times in two successive years,<br />

were labeled as residents, and comprised<br />

30% of the identified population. The abundance<br />

estimate derived by Chao’s closed<br />

model M th was 122 (95% CI = 114-140)<br />

(Chao et al. 1992) and was quite similar to<br />

the number of identified dolphins (n = 115).<br />

Taken together, the discovery rate for<br />

new individuals, the high study effort<br />

(number of surveys) extending over a 3-yr<br />

period, the high proportion of individuals<br />

showing low sighting frequencies (vis-à-vis<br />

the high number of surveys), the small<br />

number of individuals showing evidence<br />

for site fidelity (“residents”), and the relatively<br />

low abundance estimate, suggests<br />

that, as at Turneffe, a small and finite population<br />

of dolphins uses this study area.<br />

These same data indicate that the Drowned<br />

Cayes area can only support a small number<br />

of dolphins, probably due to low prey<br />

item densities (group size interpretation<br />

suggested above). Some dolphins were<br />

seen throughout and across years, however,<br />

indicating that a degree of fidelity to<br />

the area does exist.<br />

Photographs of the 115 Drowned Cayes<br />

dolphins were compared to those of 81 individuals<br />

photographed during 549 surveys<br />

at Turneffe Atoll from 1992-1996<br />

(Campbell et al. 2002). Despite the close<br />

geographic proximity of the Drowned<br />

Cayes and Turneffe, the high survey effort<br />

in both studies, and the presence of a large<br />

number of individuals in each population<br />

not considered “residents,” there was no<br />

overlap documented between these populations.<br />

While the depth of the channel<br />

separating Turneffe Atoll from the Belize<br />

Barrier Reef and the Drowned Cayes (range<br />

= 274-305 m) may act as a physical barrier<br />

to regulate movements of dolphins between<br />

the study areas, the distance between<br />

Turneffe and the Drowned Cayes<br />

could be easily traveled by this species<br />

(Stoddart 1962; Tanaka 1987; Wood 1998;<br />

Defran et al. 1999). A tentative hypothesis,<br />

NOTES <strong>17</strong>5<br />

consistent with the existing photo-identification<br />

data, is that bottlenose dolphins in<br />

Belize may have overlapping ranges along<br />

the mainland coastline; however, exploitation<br />

of the offshore islands, such as the<br />

Drowned Cayes and Turneffe Atoll, is selective<br />

among subsets of this population.<br />

Similar selective partitioning among areas<br />

of their range were shown for bottlenose<br />

dolphins (Maze and Würsig 1999) as well<br />

as humpback (Megaptera novaeangilae) and<br />

sperm (Physeter macrocephalus) whales<br />

(Clapham 2000; Whitehead and Weilgart<br />

2000). To date, no mainland coastal photoidentification<br />

surveys have been carried out<br />

in Belize. Such surveys are needed, however,<br />

in order to clarify the home ranges of<br />

Drowned Cayes and Turneffee Atoll dolphins,<br />

as well as to evaluate the selective partitioning<br />

hypothesis we have proposed.<br />

Belize is rich in natural resources that<br />

could be vulnerable to ecologically unsustainable<br />

growth in tourism, fishing, and development.<br />

The accessibility of the<br />

Drowned Cayes to the coastal mainland of<br />

Belize, including Belize City and the Belize<br />

River, places an even higher pressure on<br />

levels of resource extraction, pollution, and<br />

boat traffic in this region than in other offshore<br />

locations such as Turneffe Atoll. Although<br />

fishing pressure in the Drowned<br />

Cayes, and the rest of Belize, may be considered<br />

low compared to highly populated<br />

areas in the Caribbean, the demands on<br />

fisheries resources are increasing and the<br />

effects of overfishing are already evident as<br />

changes in fish community structure (Sedberry<br />

et al. 1999). Some suggest that reefs in<br />

the Caribbean have been overfished for<br />

centuries, and the resulting changes in<br />

community structure have placed these<br />

ecosystems in a precarious state that is now<br />

collapsing (Jackson 1997; Jackson et al.<br />

2001). While fishing in the study area is<br />

mainly artisinal, even this type of fishing<br />

may be unsustainable (Coblentz 1997; Sedberry<br />

et al. 1999). Shifts in the ecological<br />

dominance from coral to fleshy algae have<br />

been found at some reef locations in Belize<br />

(Aronson and Precht 1997; McClanahan<br />

and Muthiga 1998). Since the first major<br />

bleaching event in the history of Belize’s<br />

coral reefs occurred in 1995, an even more


<strong>17</strong>6<br />

severe event occurred in 1998 (Aronson et<br />

al. 2000). Threats, such as overfishing, pollution,<br />

and the results of climatic change,<br />

have been shown to detrimentally affect<br />

marine mammals in a variety of ways (Harvell<br />

et al. 1999; Allen and Read 2000; Fair<br />

and Becker 2000; Wilson et al. 2000; Bossart<br />

et al. 2003). In the aggregate, these factors<br />

indicate that bottlenose dolphins in Belize<br />

could play a valuable role as a sentry species<br />

in this area. For example, shifts in<br />

bottlenose dolphin occurrence and abundance<br />

might indicate a further decline in<br />

prey (and other fish species) abundance.<br />

Similarly, increases in skin lesions (observed<br />

on at least two dolphins in the<br />

Drowned Cayes population, personal observations),<br />

may provide an early signal of<br />

compromised immunity to disease (Wilson<br />

et al. 2000; Bossart et al. 2003).<br />

Acknowledgments.—We thank the following<br />

individuals and institutions for their<br />

contributions to this research: B. Winning,<br />

Oceanic Society; Oceanic Society volunteers,<br />

the Belize Fisheries Department, the<br />

Belize Forestry Department, Coastal Zone<br />

Management Authority Institute and S.<br />

Turton in Belize; E. Requeña and staff at<br />

Spanish Bay Resort, K. Schafer, K. LaCommare,<br />

A. Sanders, B. Bilgre, S. Barclay, H.<br />

Petersen, and C. Sullivan for field support;<br />

B. Bilgre, L. Hinderstein and A. Sanders for<br />

contributions to the photographic dataset;<br />

K. Dudzik, B. Hancock, J. Schivone, L.<br />

Sousa, and A. Toperoff, as well as numerous<br />

interns, at the Cetacean Behavior Laboratory,<br />

San Diego State University, for assistance<br />

with photographic analysis. We<br />

appreciate the contributions of E. Bardin,<br />

A. Bradford, B. Hancock, D. Herzing, A.<br />

Mignucci-Giannoni and one anonymous<br />

reviewer who graciously offered helpful<br />

editorial remarks on earlier drafts of this<br />

manuscript. This research was conducted<br />

under a marine scientific research permit<br />

issued to Oceanic Society from the Ministry<br />

of Natural Resources, the Environment and<br />

Industry, Government of Belize.<br />

LITERATURE CITED<br />

Allen, M. C., and A. J. Read. 2000. Habitat selection of<br />

foraging bottlenose dolphins in relation to boat<br />

NOTES<br />

density near Clearwater, Florida. Mar. Mamm. Sci.<br />

16:815-824.<br />

Aronson, R. B., and W. F. Precht. 1997. Stasis, biological<br />

disturbance, and community structure of a Holocene<br />

coral reef. Paleobiology 23:336-346.<br />

Aronson, R. B., W. F. Precht, I. G. Macintyre, and T. J.<br />

Murdoch. 2000. Coral bleach-out in Belize. Nature<br />

405:36.<br />

Bossart, G. D., et al. 2003. Pathologic findings in<br />

stranded Atlantic bottlenose dolphins (Tursiops<br />

Truncatus) from the Indian River Lagoon, Florida.<br />

Florida Scient. 66:226-238.<br />

Campbell, G. S., B. A. Bilgre, and R. H. Defran. 2002.<br />

Bottlenose dolphins (Tursiops truncatus) in<br />

Turneffe Atoll, Belize: occurrence, site fidelity,<br />

group size, and abundance. Aquat. Mamm. 28:<strong>17</strong>0-<br />

180.<br />

Clapham, P. J. 2000. The humpback whale: seasonal<br />

feeding and breeding in a baleen whale. In Cetacean<br />

societies: field studies of dolphins and whales, ed. J.<br />

Mann, R. C. Conner, P. L. Tyack, and H. Whitehead,<br />

<strong>17</strong>3-198. Chicago and London: The University<br />

of Chicago Press.<br />

Chao, A., S-M. Lee, and S-L. Jeng. 1992. Estimating<br />

population size for capture-recapture data when<br />

capture probabilities vary by time and individual<br />

animal. Biometrics 48:201-216.<br />

Coblentz, B. E. 1997. Subsistence consumption of coral<br />

reef fish suggests non-sustainable extraction. Conserv.<br />

Biol. 11:559-561.<br />

Connor, R. C., R. S. Wells, J. Mann, and A. J. Read.<br />

2000. The bottlenose dolphin: social relationships<br />

in a fission-fusion society. In Cetacean societies: field<br />

studies of dolphins and whales, ed. J. Mann, R. C.<br />

Connor, P. L. Tyack, and H. Whitehead, 91-126.<br />

Chicago and London: The University of Chicago<br />

Press.<br />

Defran, R. H., D. W. Weller, D. Kelly, and M. A. Espinoza.<br />

1999. Range characteristic of Pacific Coast<br />

bottlenose dolphins in the Southern California<br />

Bight. Mar. Mamm. Sci. 15:381-393.<br />

Fair, P. A., and P. R. Becker. 2000. Review of stress in<br />

marine mammals. J. Aquat. Ecosys. Stress Recov. 7:<br />

335-354.<br />

Grigg, E., and H. Markowitz. 1997. Habitat use by<br />

bottlenose dolphins (Tursiops truncatus) at Turneffe<br />

Atoll, Belize. Aquat. Mamm. 23:163-<strong>17</strong>0.<br />

Harvell, C. D., et al. 1999. Emerging marine diseases—<br />

Climatic links and anthropogenic factors. Science<br />

285:1505-1510.<br />

Jackson, J. B. C. 1997. Reefs since Columbus. Coral<br />

Reefs 16, Suppl.: S23-S32.<br />

Jackson, J. B. C., et al. 2001. Historical overfishing and<br />

the recent collapse of coastal ecosystems. Science<br />

293:629-638.<br />

Maze, K.S., and B. Würsig. 1999. Bottlenose dolphins<br />

of San Luis Pass, Texas: Occurrence patterns, sitefidelity,<br />

and habitat use. Aquat. Mamm. 25:91-103.<br />

McClanahan, T. R., and N. A. Muthiga. 1998. An ecological<br />

shift in a remote coral atoll of Belize over 25<br />

years. Environ. Conserv. 25:122-130.


Rogers, C. A., B. J. Brunnick, D. L. Herzing, and J. D.<br />

Baldwin. 2004. The social structure of bottlenose<br />

dolphins, Tursiops truncatus, in the Bahamas. Mar.<br />

Mamm. Sci. 20:688-708.<br />

Rossbach, K. A., and D. L. Herzing. 1999. Inshore and<br />

offshore bottlenose dolphin (Tursiops truncatus)<br />

communities distinguished by association patterns<br />

near Grand Bahamas Island, Bahamas. Can. J. Zoolog.<br />

77:581-592.<br />

Sedberry, G. R., H. J. Carter, and P. A. Barrick. 1999. A<br />

comparison of fish communities between protected<br />

and unprotected areas of the Belize reef ecosystem:<br />

implications for conservation and management.<br />

Proceedings of the Gulf and Caribbean Fisheries<br />

Institute 45:95-127.<br />

Stoddart, D. R. 1962. Three Caribbean atolls: Turneffe<br />

Islands, Lighthouse Reef, and Glover’s Reef, British<br />

Honduras. Atoll Res. Bull. 87:1-147.<br />

Tanaka, S. 1987. Satellite radio tracking of bottlenose<br />

NOTES <strong>17</strong>7<br />

dolphins Tursiops truncatus. Nippon Suisan Gakkaishi<br />

53:1327-1338.<br />

Wells, R. S., and M. D. Scott. 1999. Bottlenose dolphin<br />

Tursiops truncatus. InHandbook of marine mammals:<br />

Volume VI, The second book of dolphins and porpoises,<br />

ed. S. H. Ridgeway and R. J. Harrison, 137-182. San<br />

Diego, CA Academic Press.<br />

Wilson, B., K. Grellier, P. S. Hammond, G. Brown, and<br />

P. M. Thompson. 2000. Changing occurrence of<br />

epidermal lesions in wild bottlenose dolphins.<br />

Mar. Ecol-Prog. Ser. 205:283-290.<br />

Whitehead, H., and L. Weilgart. 2000. The sperm whale:<br />

social females and roving males. In Cetacean societies:<br />

field studies of dolphins and whales, ed. J. Mann, R. C.<br />

Conner, P. L. Tyack, and H. Whitehead, 154-<strong>17</strong>2. Chicago<br />

and London: The University of Chicago Press.<br />

Wood, C. J. 1998. Movement of bottlenose dolphins<br />

around the south-west coast of Britain. J. Zool.,<br />

London 246:155-163.


Coral Reefs (1997) 16, Suppl.: S23—S32<br />

Reefs since Columbus<br />

J. B. C. Jackson<br />

Center for Tropical Paleoecology and Archeology, Smithsonian Tropical Research Institute, Apartado 2072, Balboa, Republic of Panama<br />

Accepted: <strong>17</strong> January 1997<br />

Abstract. History shows that Caribbean coastal ecosystems<br />

were severely degraded long before ecologists began<br />

to study them. Large vertebrates such as the green turtle,<br />

hawksbill turtle, manatee and extinct Caribbean monk<br />

seal were decimated by about 1800 in the central and<br />

northern Caribbean, and by 1990 elsewhere. Subsistence<br />

over-fishing subsequently decimated reef fish populations.<br />

Local fisheries accounted for a small fraction of the fish<br />

consumed on Caribbean islands by about the mid nineteenth<br />

century when human populations were less than<br />

one fifth their numbers today. Herbivores and predators<br />

were reduced to very small fishes and sea urchins by the<br />

1950s when intensive scientific investigations began. These<br />

small consumers, most notably Diadema antillarum, were<br />

apparently always very abundant; contrary to speculation<br />

that their abundance had increased many-fold due to<br />

overfishing. Studying grazing and predation on reefs<br />

today is like trying to understand the ecology of the<br />

Serengeti by studying the termites and the locusts while<br />

ignoring the elephants and the wildebeeste. Green turtles,<br />

hawksbill turtles and manatees were almost certainly<br />

comparably important keystone species on reefs and<br />

seagrass beds. Small fishes and invertebrates feed very<br />

differently from turtles and manatees and could and can<br />

not compensate for their loss, despite their great abundance<br />

long before overfishing began. Loss of megavertebrates<br />

dramatically reduced and qualitatively changed<br />

grazing and excavation of seagrasses, predation on<br />

sponges, loss of production to adjacent ecosystems, and<br />

the structure of food chains. Megavertebrates are critical<br />

for reef conservation and, unlike land, there are no coral<br />

reef livestock to take their place.<br />

Introduction<br />

The status and trends of the world’s coral reefs are highly<br />

controversial and, until very recently, have attracted far<br />

less attention than has been lavished on the decline of<br />

tropical forests. This is partly a function of our greater<br />

ignorance about coral reefs and their briefer period of<br />

study. But it is also true that coral reef ecologists have<br />

been so devoted to dissecting small-scale processes that<br />

they have not seen the reefs for the corals.<br />

Ecology is a young science, with little reliable<br />

descriptive information from before the 1920s (Elton<br />

1927; Hutchinson 1978) and virtually no time series<br />

population data extending of back for more than a century.<br />

The situation is even worse for marine communities<br />

like coral reefs because of the difficulty of making observations<br />

underwater. Thus reef ecologists have been<br />

forced to try to explain patterns of distribution and<br />

abundance exclusively in terms of the events of the<br />

past few years using ‘‘real time’’ observations and<br />

experiments. Their success is manifest in our rapidly<br />

growing understanding of how environmental variation<br />

and biological interactions can shape reef communities<br />

(Connell 1978; Hughes 1989; Knowlton et al. 1990), although<br />

the importance of events rare on the scale of<br />

human lifetimes is only beginning to be understood<br />

(Woodley et al. 1981; Jackson 1991; Knowlton 1992;<br />

Hughes 1994).<br />

In the process of these endeavors, however, reef ecologists<br />

have turned their backs on history and assumed<br />

that what they were studying was ‘‘normal,’’ despite the<br />

lack of rigorous baseline data for the condition of coral<br />

reefs preceding the industrial revolution and the onset of<br />

worldwide, exponential human population growth. As<br />

a result, many reef ecologists have concluded that coral<br />

reefs are healthy in the face of overwhelmingly increasing<br />

evidence to the contrary. Indeed, the idea for this study<br />

arose from my feeling like Cassandra in the face of such<br />

denial during the 1993 Miami Colloquium on Global<br />

Aspects of Coral Reefs: Health, Hazards and History. Most<br />

felt at the beginning of the colloquium that the condition<br />

of coral reefs was on the whole rather good, despite<br />

Wilkinson’s (1992) plenary paper at the 7th <strong>International</strong><br />

Coral Reef Symposium about widespread devastation<br />

worldwide. However, by the end of the Colloquium, there<br />

was a beginning acceptance that the situation was perhaps<br />

bad in the Caribbean, but only recently, and probably not<br />

in the pacific.


S24<br />

The problem is that everyone, scientists included, believes<br />

that the way things were when they first saw them is<br />

natural. However, modern reef ecology only began in the<br />

Caribbean, for example, in the late 1950s (Goreau 1959;<br />

Randall et al. 1961; Randall 1965) when, enormous changes<br />

in coral reef ecosystems had already occurred. The same<br />

problem now extends on an even greater scale to the<br />

SCUBA diving public, with a whole new generation of<br />

sport divers who have never seen a ‘‘healthy’’reef, even by<br />

the standards of the 1960s. Thus there is no public perception<br />

of the magnitude of our loss.<br />

Another insidious consequence of this ‘‘shifting baseline<br />

syndrome’’ (Pauly 1995; Sheppard 1995) is a growing ecomanagement<br />

culture that accepts the status quo,andfiddles<br />

with it under the mantle of experimental design and statistical<br />

rigor, without any clear frame of reference of what it is<br />

they are trying to manage or conserve. These are the coral<br />

reef equivalents of European ‘‘hedgerow ecologists’’ arguing<br />

about the maintenance of diversity in the remnant<br />

tangle between fields where once there was only forest.<br />

Let me say from the start that I am not going to make<br />

some romantic appeal to set back the clock, nor propose<br />

draconian scenarios that ignore the realities of inexorable<br />

human population growth and underdevelopment. Instead,<br />

my goal is to set the stark realities we face in a deeper<br />

historical perspective than the last few decades in order to<br />

(1) silence absurd notions of multiuse sustainability and (2)<br />

help to define better the limited alternatives available. I will<br />

limit my discussion to the Caribbean, and principally to<br />

Jamaica, because I know the Caribbean best and Jamaica is<br />

arguably the worst case today in that region. Jamaica also<br />

offers the best historical record, which extends back nearly<br />

350 years, and provides startlingly good but unexploited<br />

information for ecological assessment.<br />

I will first work backward from the present to re-examine<br />

aspects of the classic story relating overfishing, the mass<br />

mortality of the long-spined sea urchin Diadema antillarum,<br />

and the collapse of Caribbean coral reefs in Jamaica and<br />

elsewhere (Lessios et al. 1984; Hay 1984; Hughes 1994). This<br />

is not to deny the important consequences of overfishing,<br />

but to show that we have been at least partly wrong about<br />

the historical role of the sea urchin. I will then work<br />

forward from 1492 to examine the extraordinarily rapid<br />

depletion of large consumers on coral reefs and their environs,<br />

which were once the equivalents of the wildebeeste<br />

and elephants of the Serengeti plains (Sinclair and Arcese<br />

1995). For this purpose, I will emphasize the Jamaican<br />

basedfisheryofthegreenturtleChelonia mydas because the<br />

data are the best; but the same sort of story applies to<br />

sharks, rays, groupers, manatees and the extinct Caribbean<br />

monk seal. I will then depart from the general theme of<br />

overfishing to briefly consider inputs from the land. These<br />

have almost certainly been of comparable importance to<br />

overfishing but the data are less complete. Finally, I will<br />

return to Jamaica to examine the fishing history there since<br />

Columbus, which clearly shows that coral reef ecosystems<br />

had begun to fall apart in the eighteenth century.<br />

Diadema, damselfish and overfishing<br />

The conventional story (Hay 1984; Lessios 1988; Hughes<br />

1994) is that intense overfishing allowed Diadema<br />

to increase because of reduced predation by fishes, and<br />

competitive release for algal food that was no longer<br />

consumed by larger herbivores. This increase in Diadema<br />

compensated for the loss of herbivorous fishes and kept<br />

down the growth of seaweeds on reefs. Then when the<br />

Diadema suddenly died, there were no other consumers<br />

capable of cropping the seaweeds which soon overgrew<br />

and killed most of the reef corals at depths down to about<br />

50 m.<br />

What are the facts and what is the inference in this<br />

story? The facts are that (1) overfishing was extreme by<br />

any standard, (2) Diadema was extremely abundant before<br />

the mass mortality in 1983, (3) mass mortality of Diadema<br />

allowed a dramatic increase in seaweeds which overgrew<br />

and smothered corals. The inference is that (1) decrease in<br />

large fish allowed a large increase in Diadema, (2) increase<br />

in people caused the decrease in fish in a roughly proportional<br />

fashion, and (3) the effects of people were relatively<br />

recent. Regarding the latter point, for example, Hughes’<br />

(1994) graph of human population increase starts in 1850<br />

and is introduced within a section titled ‘‘Overfishing:<br />

1960s to Present’’.<br />

Let us examine some problems with this inference<br />

before reviewing the historical data. Hay’s (1984) pioneering<br />

regional study of Diadema versus fish grazing confounds<br />

‘‘overfished’’ and ‘‘less fished’’ with geography. Hay<br />

was very careful to avoid terms like ‘‘pristine’’ and ‘‘unfished’’,<br />

unlike many who have referred to his work subsequently.<br />

However, all of his ‘‘overfished’’ reefs are on<br />

islands in the central Caribbean, and all but one of his<br />

‘‘less fished’’ reefs are on the mainland. Moreover, Levitan’s<br />

(1992) ingenious study of Diadema nutrition over the<br />

last century shows an even greater geographic effect. Levitan<br />

(1991) showed experimentally that the ratio of the size<br />

of the feeding apparatus to the size of the animal test<br />

increases in inverse proportion to the food supply. He<br />

then examined changes in the ratio through time using<br />

museum specimens, and found a significant but surprisingly<br />

small increase over the past century when Diadema<br />

populations were inferred to have exploded due to overfishing.<br />

In contrast, geographic variation accounted for<br />

much more (23%) of the total variation observed than<br />

that due to time. Ordination of coral abundance data<br />

from reefs around the Caribbean compiled by Liddell and<br />

Ohlhorst (1988) also shows strong island-mainland differences<br />

in overall reef community structure (Jackson et al.<br />

1996).<br />

So is it possible that Diadema were abundant in the<br />

Caribbean before Columbus arrived? It turns out that<br />

Diadema has long attracted commentary because of its<br />

great abundance and reputation of being dangerous<br />

(Table 1). Moreover, all of the authors cited except Young<br />

were professional or amateur naturalists, with extensive<br />

experience dredging or (in the case of Beebe 1928) diving<br />

in tropical waters. All were well known for the reliability<br />

and accuracy of their observations and were not the type<br />

to mistake an occasional aggregation or hearsay for genuinely<br />

great abundance. These sources make it clear that<br />

Diadema was indeed very abundant in the seventeenth<br />

century when human populations were very small, and<br />

therefore long before overfishing could have caused their<br />

increase.


Table 1. Caribbean Diadema lore<br />

Source Location Quoted text<br />

Beebe 1928 Haiti Under every bit of coral 2 in great abundance<br />

Clark 1919 Jamaica and Dry<br />

Tortugas<br />

Nutting 1919 Barbados Antigua<br />

and Bahamas<br />

On and about the coral reefs, the dreaded poisonous ‘‘black<br />

sea egg’’ (Centrechinus antillarum) is common and on certain<br />

areas it is so numerous that a person can scarcely move<br />

about without touching one.<br />

No one goes bathing or into the water for any purpose in<br />

this region without being warned against the danger of being<br />

wounded by the cruel black spines of this ubiquitous<br />

sea-urchin. It is found almost everywhere in shallow water,<br />

both on sandy and rocky bottom.<br />

The all too familiar black sea-egg Diadema antillarum is<br />

abundant here, as it is everywhere that I have collected in the<br />

West Indies (therefore includes his 1895 expedition to the<br />

Bahamas, not seen)<br />

Henderson 1914 Cuba We were then upon the inner edge of the main reef upon<br />

which any further progress would have been difficult on<br />

account of the rapidly increasing numbers of the long<br />

black-spined sea urchins, the diademas2<br />

2the usual presence of the net [dredging] of the diadema<br />

sea-urchin2<br />

During the hours of bright sunshine the diademas seek cover<br />

under the rocks2 In the late afternoon2they issue forth<br />

en masse in search of food and probably continue their slow<br />

wanderings throughout the night. In localities where hiding<br />

places are few, such as upon sandy patches in or near a reef,<br />

the diademas are always more or less in evidence.<br />

Field 1891 Jamaica The most strikingly conspicuous of all the creatures about<br />

the coral reefs. As one approaches the cay, far down in the<br />

water are seen numerous irregularly shaped black patches of<br />

varying extent.<br />

2around and under its branches (elkhorn coral) are<br />

crowded together great numbers of this large black urchin.<br />

2what formidable defense against attack they present<br />

when crowded so closely together.<br />

Aggasiz 1883 Dry Tortugas 2on somewhat deeper regions (of the reef) we find pockets<br />

filled with large Diadematidae.<br />

Young 1847<br />

(writing of his<br />

observations<br />

in 1839)<br />

Sloane <strong>17</strong>25<br />

(writing of his<br />

observations<br />

in 1688)<br />

Nicaragua and<br />

Honduras<br />

There is also strong paleontological evidence that Diadema<br />

were abundant on Jamaican reefs long before any<br />

humans arrived there (Gordon and Donovan 1992). Diadematoid<br />

plates and spines are the most abundant<br />

echinoid remains in the 125 000 year old Falmouth<br />

Formation, constituting 90% of all identified echinoid<br />

fragments. These are almost, certainly of Diadema antillarum<br />

because the only other Caribbean Pleistocene diadematoid<br />

is Astropyga magnifica which occurs today in<br />

S25<br />

2they were both badly cut by the coral and sea eggs in<br />

diving for the things that had been upset.<br />

Numbers of sea eggs were seen in all directions, and we well<br />

knew the danger of getting amongst them, as they have long<br />

and sharp pointed spines, which inflict deep and dangerous<br />

wounds on those who chance to tread on them.<br />

2the handsome sea-eggs inviting but to betray2<br />

Jamaica The great, long prickled Sea Egg2set about on every hand<br />

with prickles, the largest being three or four inches long, with<br />

membrane round their setting on to the shell2purple deep<br />

coloured2<br />

The prickles of this Sea Echinus are very rough and considered<br />

poisonous.<br />

I have found them in great numbers on the reef by Gun-Key,<br />

or, Cayos off the Port Royal Harbour in great numbers.<br />

deeper water than the backreef environment of the Falmouth<br />

Formation (Donovan and Gordon 1993).<br />

In conclusion, Diadema apparently has been the most<br />

abundant sea urchin on Caribbean reefs for at least<br />

125 000 years. Its abundance still may have increased<br />

historically due to overfishing, but we lack the quantitative<br />

data to tell. However, it now seems unlikely that any<br />

such increase was as great as that, for example, of<br />

Echinometra mathaei in response to over-fishing in


S26<br />

Kenyan lagoons (McClanahan and Muthiga 1988;<br />

McClanahan et al. 1996). Thus Diadema abundance is at<br />

best a secondary consequence of the degradation of Caribbean<br />

reefs, and we need to look for other indicators of<br />

faunal change in response to human interference. For this<br />

purpose, let us now turn to changes in the abundance of<br />

the really large consumers in reef environments, such as<br />

green turtles.<br />

How many turtles in 1492?<br />

Big animals are ecologically important, not only because<br />

of the amount of plants or animals each individual consumes,<br />

but also because of the physical and biological<br />

disturbance they cause, which fundamentally alters their<br />

environment and affects other species. Perhaps the best<br />

studied example is the Serengeti ecosystem of east Africa<br />

(Sinclair 1995a; Sinclair and Arcese 1995). Long distance<br />

migration (wildebeeste) and growth to very large size<br />

(elephant, hippopotamus, rhinoceros and buffalo) result in<br />

virtual escape from predation, so that such herbivores are<br />

‘‘bottom up’’ regulated by food limitation rather than ‘‘top<br />

down’’ by predators (Sinclair 1995a). The enormously<br />

abundant wildebeeste is a keystone species because its<br />

grazing and migration directly or indirectly affect almost<br />

everything else. Wildebeeste grazing alters the protein<br />

content of the grass and stimulates growth of new shoots,<br />

increasing the available food supply for smaller grazers,<br />

and possibly also for themselves (McNaughton and Banyikwa<br />

1995). There is also strong evidence for alternate<br />

stable states of vegetation (woodland versus grassland),<br />

maintained by grazing and disturbance by elephants and<br />

by fire (Dublin 1995; Sinclair 1995b).<br />

None of these topics have been studied as well on coral<br />

reefs and surrounding seagrass environments, but what we<br />

do know strongly suggests similarly important ecosystem<br />

effects of large species (Heinsohn et al. 1977; Ogden 1980;<br />

Thayer et al. 1984; Lanyon et al. 1989; Sheppard et al.<br />

1992). Green turtles crop ‘‘turtlegrass’’ ¹halassia testudinum<br />

to only a few centimeters above the bottom, and<br />

cause erosion and pits in the rhizomal mat of turtlegrass<br />

beds. Manatees and dugongs do much the same, sometimes<br />

ripping up entire beds of seagrasses and other<br />

aquatic vegetation, and thereby causing even greater<br />

physical disturbance. Stingrays excavate pits in seagrass<br />

beds foraging for mollusks beneath the rhizome mat,<br />

hawksbill turtles rip up sponges, and large parrotfishes<br />

occasionally bite to pieces entire coral colonies for unknown<br />

reasons.<br />

I have nothing new to add to such observations, except<br />

the comment that I have not even seen most of these large<br />

animals underwater for twenty years or more, and some of<br />

them never at all, despite thousands of hours SCUBA<br />

diving on and around coral reefs. Instead, I want to dwell<br />

on the past enormity of the populations of such creatures<br />

using green turtles as an example. My calculations are<br />

rough in ways that responsible turtle biologists have shied<br />

away from, and the data are 15 to 300 years old. Nevertheless,<br />

it is essential to try, because we have no conception of<br />

the way things were. Why, after all, are so many hundreds<br />

of sites around the Caribbean, such as the Dry ¹ortugas,<br />

named after turtles that almost no living person has<br />

ever seen?<br />

Calculations based on old hunting data from the Cayman<br />

Islands<br />

Estimates of the size of pre-columbian human populations<br />

in the Caribbean are controversial and depend on differing<br />

interpretations of archeological and historical sources<br />

(Roberts 1989). Nevertheless, populations of Hispaniola,<br />

Jamaica and Cuba certainly ranged in the hundreds of<br />

thousands, perhaps even in the millions, and were sustained<br />

by a highly productive agricultural system supplemented<br />

by fishing and hunting (Sauer 1966; Rouse 1992).<br />

These early Americans were reduced by conquest, slavery<br />

and disease to only a few thousand by 1600. Subsequent<br />

Spanish colonization was slow, so that there were only<br />

about five thousand people in Jamaica when the English<br />

captured the island in 1655 (Long <strong>17</strong>74). There was also<br />

no effective agricultural base, so the English turned immediately<br />

for food to the vast populations of green turtles<br />

that nested on Grand Cayman Island (Table 2). These<br />

abundant turtles were essential to the rapid growth and<br />

success of Jamaica as England’s most important colony of<br />

the time, and indeed provided most of the meat consumed<br />

there until the <strong>17</strong>30s (Sloane <strong>17</strong>07, <strong>17</strong>25; Long <strong>17</strong>74; Lewis<br />

1940; King 1982).<br />

Thirty years later, the Cayman Islands fishery had<br />

grown to approximately 40 sloops and 120 to 150 men<br />

who brought back to Jamaica some 13 000 turtles per year<br />

between 1688 and <strong>17</strong>30. Let us assume that the sex ratio<br />

and migration interval (time between years that females<br />

reproduce) were 1: 1 and 2.5 years respectively, just as they<br />

are today (Bjorndal 1982). Thus, the proportion of the<br />

adult population (N ) that are nesting females (N )is<br />

<br />

given by<br />

N "N /0.5 (sex ratio)0.4 (migration interval)"5N .<br />

<br />

Let us further assume that hunters in 1688 captured only<br />

1% of the nesting female turtles per year. This is an<br />

arbitrary but almost certainly conservative guess for the<br />

purpose of illustration, based on the impression from the<br />

early descriptions that the beaches all around Grand<br />

Cayman were literally covered by turtles, so that 13 000<br />

would have been a very small fraction of the total. Moreover,<br />

female green turtles require 40—60 years to reach<br />

reproductive maturity (Bjorndal and Zug 1995), so that<br />

harvested females could not have been replaced for half<br />

the century-long duration of the fishery. Despite this,<br />

however, 13 000 reproductively mature females were harvested<br />

annually over 42 y, for a total (with the above<br />

correction) of more than 2.5 million on this basis alone.<br />

Based on the assumption of an initial catch rate of 1%, the<br />

estimated total adult population (N ) based on the early<br />

<br />

hunting data is<br />

N "513 000 (number harvested)/0.01 (% N caught)<br />

<br />

"6.5 million.<br />

Further assume that there were five additional precolumbian<br />

green turtle rookeries roughly equal to Grand Cay-


Table 2. Historical accounts of<br />

the early great abundance of<br />

green turtles in the Caribbean<br />

Andres Bernaldez,<br />

writing about<br />

Columbus’<br />

2nd voyage in 1494<br />

Ferdinand Columbus,<br />

writing about the 4th<br />

voyage in 1503<br />

Edward Long (<strong>17</strong>74),<br />

writing of the late<br />

1600s<br />

not seen, cited in Lewis 1940<br />

Southeastern<br />

Cuba<br />

Cayman<br />

Islands<br />

man including Bermuda, Bahamas, Florida Keys,<br />

Costa Rica, and Isla Aves (now less than a mile long but<br />

historically much larger). Then the estimated total adult<br />

population for the entire precolumbian Caribbean is five<br />

to six times the Grand Cayman estimate, or about 33 to 39<br />

million. This is about 15 to 20 times the abundance and<br />

biomass of large ungulates in the Serengeti today (Sinclair<br />

1995a)!<br />

Calculations based on carrying capacity<br />

The following is based on Bjorndal’s (1982) study of green<br />

turtle nutrition and life history. There is apparently no<br />

reliable compilation of the total area of seagrasses in the<br />

wider Caribbean. However, we can assume that roughly<br />

ten percent of the total shelf area, which is 660 000 km<br />

excluding south Florida (Munroe 1983), is covered by<br />

seagrasses for a total of 66 000 km. This is probably<br />

conservative, since the mapped area of seagrasses for<br />

south Florida alone is 5500 km (Ziemann 1982). Bjorndal<br />

(1982) calculated that the carrying capacity of closely<br />

cropped (2.5 cm) ¹halassia is one 100 kg adult female per<br />

72 m per year, which rounding up to one turtle per<br />

100 m, gives 10 000 adult females per km. Assume further<br />

that the carrying capacity is the same for males as for<br />

females. Then the estimated total adult population (N )<br />

for the entire Caribbean is<br />

N "10 00066 000"660 million,<br />

which is about 20 times the estimate based on the old<br />

hunting data. Of course, sharks and other predators of<br />

principally juvenile turtles were also very much more<br />

abundant. However, Sinclair’s (1995a) generalization<br />

about the predominantly ‘‘bottom up’’ regulation of large,<br />

migratory herbivores suggests that the abundance of<br />

large, migrating adult green turtles would have approached<br />

carrying capacity even in the face of intense<br />

predation on juveniles.<br />

Differences in feeding between green turtles and other<br />

herbivores<br />

One adult green turtle consumes roughly the same<br />

amount of turtlegrass as 500 large sea urchins like Dia-<br />

West of the<br />

Cayman<br />

Islands<br />

S27<br />

But in those twenty leagues, they saw very many more, for<br />

the sea was thick with them, and they were of the very<br />

largest, so numerous that it seemed that the ships would run<br />

aground on them and were as if bathing in them.<br />

2in sight of two very small and low islands, full of tortoises,<br />

as was all the sea about, insomuch that they looked like little<br />

rocks2<br />

2it is affirmed, that vessels, which have lost their latitude in<br />

hazy weather, have steered entirely by the noise which these<br />

creatures make in swimming, to attain the Cayman isles.<br />

dema antillarum or ¹ripneustes ventricosus, which works<br />

out to a potential increase of about 5 sea urchins per m of<br />

seagrass beds throughout the Caribbean (calculations<br />

based on data in Thayer et al. 1984). Much more significantly,<br />

however, there are profound differences in the<br />

ways turtles versus sea urchins and herbivorous fishes<br />

graze on turtlegrass, and how well they process what they<br />

eat, with important consequences for the structure and<br />

function of the entire turtlegrass ecosystem (Ogden 1980;<br />

Thayer et al. 1982, 1984; Ogden et al. 1983). Sea urchins<br />

and fishes tend to feed indiscriminantly on turtlegrass, or<br />

on the older parts of the blades, whereas green turtles crop<br />

blades close to their base. They also return repeatedly to<br />

the same discrete grazing plots which may be maintained<br />

for a year or more. Moreover, grazing sea urchins and<br />

fishes feed principally on the cell contents of seagrass<br />

blades, due to lack of appropriate enzymes or microflora<br />

to digest cell walls, whereas green turtles rely on microbial<br />

fermentation in the hindgut to digest cell walls as well as<br />

their contents (Thayer et al. 1984).<br />

Repeated grazing of the same plots of turtlegrass by<br />

green turtles temporarily increases the nutritional quality<br />

of the blades for the turtles (Thayer et al. 1984). However,<br />

it also stresses the plants and eventually reduces turtlegrass<br />

productivity, when turtles presumably move on to<br />

feed elsewhere. Turtle grazing also results in a roughly<br />

15-fold decrease in the supply of nitrogen to seagrass roots<br />

and rhizomes, due to greatly decreased accumulation of<br />

detritus and digestion of cell walls (Thayer et al. 1984).<br />

Nitrogen in turtle feces and urine is also released over<br />

a much wider area with resulting net export to adjacent<br />

coral reefs and other adjacent ecosystems.<br />

These differences in the effects of grazing by small and<br />

large herbivores extend to dugongs and manatees that<br />

also possess hindgut microflora that digest cell walls<br />

(Thayer et al. 1984), and were also formerly very abundant<br />

throughout tropical seas (Dampier <strong>17</strong>29, Sheppard et al.<br />

1992). Similar ecosystem effects almost certainly transpired<br />

on coral reefs due to the virtual disappearance of<br />

large predators such as hawksbill turtles, groupers, and<br />

sharks that were also extremely abundant historically<br />

(Ibid, King 1982; Limpus 1995). For example, the hawksbill<br />

turtle, Eretmochelys imbricata, feeds almost exclusively<br />

on sponges which commonly display large, characteristic<br />

feeding scars in areas where the turtles are still common<br />

(Meylan 1985, 1988; van Dam and Diez, in press). Hawksbills<br />

can rip big sponges apart, and in the process facilitate


S28<br />

predation by other sponge feeders that cannot penetrate<br />

the heavy armor of many sponges such as Geodia; unlike<br />

the much smaller angelfishes that also feed almost exclusively<br />

on reef sponges (Randall and Hartman 1968). Hawksbills<br />

feed today mostly on non-toxic astrophorid and<br />

hadromerid sponges, but this may not have been true in<br />

the past when, by analogy to green turtles (King 1982;<br />

Limpus 1995; this study), hawksbills almost certainly<br />

numbered in the tens of millions. Thus non-toxic sponges<br />

may have been proportionally rarer before hawksbills<br />

were intensely harvested.<br />

Large herbivores and carnivores are ecologically extinct<br />

on Caribbean coral reefs and seagrass beds, where food<br />

chains are now dominated by small fishes and invertebrates<br />

(Hay 1984, 1991; Knowlton et al. 1990). Moreover,<br />

similar depletion of megavertebrates is almost complete<br />

throughout the Indo-Pacific (Sheppard et al. 1992; Limpus<br />

1995). Small consumers cannot fully compensate for<br />

the loss of megavertebrates because they cannot capture,<br />

consume or process their prey in the same ways as larger<br />

species. Many small herbivores are also feeding specialists,<br />

and live commonly on prey that are chemically defended<br />

against the much larger consumers that have now disappeared<br />

(Hay 1991, in press). Thus coral reef ecosystems<br />

must function in fundamentally different ways than only<br />

a few centuries ago. Similarly great changes are going on<br />

right now in east Africa (Sinclair and Arcese 1995). They<br />

also occurred 10 000 years ago in neotropical forests when<br />

over 15 genera of large herbivores became extinct (Janzen<br />

and Martin 1982), before which forests were probably<br />

more of a mixture of open forest and grassland than the<br />

dense tropical forest we imagine as more natural (Janzen<br />

and Wilson 1983). As a result, neotropical herbivorous<br />

food chains are now dominated by insects and small<br />

mammals, except where free-ranging livestock may have<br />

partially redressed the balance. But there are no such<br />

livestock on coral reefs!<br />

Effects from the land<br />

Sedimentation caused by deforestation and poor agricultural<br />

practice, eutrophication, and oil pollution have<br />

greatly increased along Caribbean coasts during the<br />

past few decades (Rodriguez 1981; Lugo et al. 1981),<br />

causing widespread and dramatic decline of coral reefs<br />

and associated marine communities throughout the<br />

region (Cortés and Risk 1985; Rogers 1985, 1990;<br />

Tomascik and Sander 1985, 1987a, b; Bak 1987; Jackson<br />

et al. 1989; Guzmán et al. 1991). A common assumption in<br />

such studies is that most reefs were not seriously affected<br />

by runoff from the land before the observations began.<br />

This allows the investigator to designate ‘‘unaffected’’ reefs<br />

or corals that can be used as a baseline to measure the<br />

effects of a particular source of pollution, such as an oil<br />

spill.<br />

New evidence, however, suggests that great ecological<br />

changes due to runoff began long before modern ecological<br />

analyses of Caribbean reefs. For example, Montastrea<br />

‘‘annularis’’ and Porites spp. were the dominant reef building<br />

corals around Barbados in the 1960s and 1970s (Lewis<br />

1960; Macintyre 1968; Stearn et al. 1977). These began to<br />

decline dramatically in the 1970s and 1980s, due primarily<br />

to runoff and eutrophication caused by exponential increase<br />

in populations of residents and tourists (Tomascik<br />

and Sander 1985, 1987a, b; Bell and Tomascik 1993).<br />

However, extensive deforestation began in Barbados in<br />

1627 for sugar plantations, after which the vegetation of<br />

the island was completely destroyed three times by hurricanes<br />

(references in Lewis 1984); with untold increases in<br />

runoff of sediments from the land. Moreover, shallow reefs<br />

at that time were dominated by the elkhorn coral Acropora<br />

palmata, which persisted in huge tracts all along the<br />

southern and western coasts of the island until the 1920s<br />

(Nutting 1919; Lewis 1984; Bell and Tomascik 1993), and<br />

the same was true throughout the Late Pleistocene (Mesolella<br />

1957; Jackson 1992). Degradation of these reefs and<br />

seagrass beds was clearly visible in aerial photographs<br />

taken in the 1950s (Lewsey 1978) when elkhorn corals<br />

were rare (Lewis 1960).<br />

Similar decline in Acropora palmata and A. cervicornis<br />

occurred all along the coast of lower Central America in<br />

the 1970s and 1980s due to disease, deforestation, algal<br />

overgrowth due to the decline of Diadema, coral bleaching,<br />

oil spills, and other factors (Cortés and Risk 1985;<br />

Cortés 1993; Guzmán et al. 1991; Ogden and Ogden 1993).<br />

Live coral cover along the Caribbean coast of central<br />

Panama has declined by 50—90% in the past ten years<br />

(Guzmán and Jackson, unpublished data). However, there<br />

were hidden signs of danger long before, as measured by<br />

a steady decline of nearly 50% in the growth rate of the<br />

massive coral Siderastrea siderea over the past century<br />

(H. Guzmán, unpublished data). This is all the more<br />

remarkable because S. siderea generally prospers in turbid<br />

coastal environments unsuitable to most other Caribbean<br />

coral species. Guzmán’s work obviously needs to be replicated<br />

elsewhere, but it strongly suggests that reef environments<br />

had begun to deteriorate at least 100 years<br />

before coral cover began to seriously decline. Isotopic<br />

ratios provide excellent proxy records of climate change,<br />

but simple growth rates and incidence of injuries are<br />

measures of coral fitness through time (Jackson 1982). As<br />

such, they may be among the best available measures of<br />

coral reef health (Dodge and Vaisnys 1977; Guzmán et al.<br />

1994; Jackson 1995).<br />

Fishing and people in Jamaica and San Blas<br />

The green turtle fishery in the Cayman Islands crashed in<br />

the latter half of the eighteenth century, and was entirely<br />

gone by 1800 when the Cayman islanders moved on to do<br />

the same thing to the turtles of the Moskito Coast (Long<br />

<strong>17</strong>74; Lewis 1940; Carr 1956; Neitschmann 1973, 1982;<br />

King 1982). Fishing on Jamaican reefs was inadequate to<br />

make up for the loss of turtles. By 1881 locally caught fish<br />

accounted for only 15% of the total consumed in Jamaica,<br />

with imported dried and preserved fish from the temperate<br />

zone making up the balance (Duerden 1901). This was<br />

probably true by the early 1800s, but there are not enough<br />

quantitative data. Moreover, extensive trials had clearly<br />

demonstrated that there was little prospect for improvement<br />

of local fisheries by trawling or longline fishing which<br />

are unsuitable for areas of coral reefs (Duerden 1901).


Table 3. Excerpts from Ernest F.<br />

Thompson’s 1945 report ¹he<br />

fisheries of Jamaica<br />

Despite these realities, Duerden’s (1901) official report<br />

was surprisingly optimistic and, in a still all too familiar<br />

tone, called for more ‘‘scientific investigations and encouragement’’<br />

to improve the marine resources of the Caribbean<br />

region. Half a century later the fisheries of Jamaica<br />

had not improved and were clearly unimprovable<br />

(Thompson 1945, Table 3). Thompson was far ahead of his<br />

time in recognizing the need for greatly reducing the<br />

numbers of fishermen, helping them to find alternative<br />

livelihoods, and focusing instead on the economic opportunities<br />

of fishing for tourism. But his report was apparently<br />

ignored and overfishing continued unabated.<br />

By far the most extensive coral reef fisheries research<br />

project in the Caribbean was carried out from 1969—1973<br />

all around Jamaica and in the Pedro Cays (Munroe 1983).<br />

Overfishing was accepted as an established fact, and it was<br />

‘‘shown that for most areas of the Jamaican Shelf, the<br />

fishing intensity is sufficient to ensure that extremely few<br />

fishes survive for more than a year after recruitment, and<br />

the proportion of fishes which survive to spawn must be<br />

extremely small.’’ The report goes on to recommend<br />

a mesh size for fish traps of 6.60 cm maximum aperture<br />

that would provide a maximum yield of barely reproductive<br />

juveniles with a maximum fishing intensity of 1.5<br />

canoes/km, without any consideration of the possible<br />

consequences for the health of the entire coral reef ecosystem.<br />

Of course, it is easy to be unfairly critical with the<br />

hindsight of the Diadema debacle and the collapse of<br />

Jamaican reefs (Hughes 1994), although even in the 1970s<br />

the consequences of not having any pretty reef fishes<br />

larger than sardines on lost tourist revenues were clear.<br />

The situation in Jamaica is a story repeated everywhere<br />

throughout the Caribbean, including the traditional fisheries<br />

of indigenous peoples commonly romanticized for exhibiting<br />

wise restraint from overharvesting not observed<br />

by others. The Comarca Kuna Yala, for example, extends<br />

some 250 km along the eastern Caribbean coast of Panamá<br />

and contains extensive coral reefs, seagrass beds<br />

and mangroves (Porter 1972; Glynn 1973; Clifton<br />

et al. 1996). The population of Kuna people within the<br />

Comarca increased from less than 9000 in 1904 to less<br />

than 24 000 in 1970, when marine biological research<br />

intensified in the Comarca, and had climbed to 41 000 in<br />

1989 (Francisco Herrera, personal communication). Kuna<br />

Locally caught fish represents less than 15% of the protein fish food consumed in Jamaica2 There is<br />

little prospect of any large increase in this local catch. In fact the probability is that the local areas are<br />

already overfished (p. 5)<br />

2the greatest need for Jamaican fishermen is more in the nature of welfare than development (p. 7,<br />

Thompson’s italics).<br />

In a great many countries, fishing has proved a very valuable asset as a tourist attraction. Development<br />

of the tourist potential would require that the present usages of some sections of the community must be<br />

controlled and restricted (p. 7).<br />

The fishing situation in Jamaica can be very briefly summarised. There are too many men trying to catch<br />

too few fish. As there seems little prospect of increasing the number of fish available, the only thing to do,<br />

if a decent living standard is to be attained, is to reduce the number of fishermen. Thus the chief problem<br />

for Jamaican fishermen is to organize them, stabilise their economy and assist about four-fifths of them<br />

to drift back to agriculture from whence they came (p. 83)<br />

S29<br />

artesanal fishing has always been a small-scale enterprise,<br />

and the Comarca is closely guarded against outside exploitation.<br />

Nevertheless, reefs were severely overfished by<br />

the 1970s when large fishes were uncommon and branching<br />

acroporid and poritid corals had been mined extensively<br />

for landfill. Thus 250 km of coast were already<br />

insufficient for 24 000 Kuna, ten years before lobster fishing<br />

for external markets reduced lobster populations to<br />

critically low levels within only ten years (Chapin 1995;<br />

Ventocilla et al. 1995).<br />

So let us now return to Jamaica to consider the history<br />

of Jamaican fisheries in light of human growth, much as<br />

Hughes (1994) did, but beginning 350 years before (Fig. 1).<br />

Depopulation by the Spaniards in the sixteenth century<br />

left the island almost uninhabited until the British invasion<br />

in 1655, when marine life may well have been at its<br />

apogee of the past 10 000 years. In 1688, when Sloane was<br />

in Jamaica, there were still only 40 000 Jamaicans and<br />

Diadema was the most abundant sea urchin on coral reefs.<br />

In <strong>17</strong>93, there were 300 000 Jamaicans and breadfruit had<br />

just been introduced following extensive research by the<br />

Royal Society and the mutiny on the Bounty, to help stave<br />

off the starvation of Jamaican slaves. In 1881, there were<br />

700 000 Jamaicans and local fish accounted for only 15%<br />

of the fish consumed. In 1945 there were 1 350 000 Jamaicans,<br />

and the 5 500 000 kg of fish caught locally was still<br />

only 15% of that consumed. In 1962 there were 1 700 000<br />

Jamaicans and the fish harvest peaked at 11 000 000 kg.<br />

This was also when Goreau (1959) published his first<br />

famous paper about Jamaican coral reefs and the modern<br />

ecological perspective was born. By 1968, when I began<br />

my own research in Jamaica, there were 1 900 000 Jamaicans,<br />

and the harvest of minnow-sized fishes from Jamaican<br />

coastal waters was back down to 5 500 000 kg.<br />

Munroe’s fisheries research project was only just beginning.<br />

It is obvious that any direct relationship between human<br />

population growth and fishing in Jamaica ended in<br />

the eighteenth century when human populations were<br />

only 10% of the present. Throughout this time, Jamaicans<br />

kept on eating fish courtesy of the now collapsed Grand<br />

Banks fisheries, and the same was true throughout the<br />

Caribbean. Thus, the causes of the present ecocatastrophe<br />

are deep and historical, not just the almost ‘‘current<br />

events’’ that have passed as history before.


S30<br />

Fig. 1. Jamaican human population growth since Columbus and the<br />

depletion of local fisheries resources. Fisheries became inadequate<br />

some time in the mid nineteenth century when the local population<br />

was about 15% of that today. Value for 1492 arbitraily set at 100 000<br />

which is almost certainly much too low (Sauer 1966). Sources for<br />

population size: 1658—<strong>17</strong>68 (Long <strong>17</strong>74), 1844—1871 (Gardner 1909).<br />

1901 (Duerden 1901), 1920—2000 (Hughes 1994), 1980 (National<br />

Geographic Society 1981)<br />

Concluding remarks<br />

I hope this brief discussion will put to rest two dangerous<br />

stupidities, at least within the scientific community. The<br />

first is the placebo of sustainable use for everyone and the<br />

second is the fallacy of a ‘‘pristine’’ coral reef. Forty<br />

thousand Kuna are too many fish eaters for 250 km of<br />

coastline (Ventocilla et al. 1995), just as a few hundred<br />

thousand Jamaicans were too many fish eaters for<br />

Jamaica. The same is true for reefs everywhere else, and<br />

not just the developing world (Wilkinson 1992). Even the<br />

Great Barrier Reef cannot be used sustainably at present<br />

levels, despite the best protection in the world (Bradbury<br />

et al. 1992), and virtually all other reefs are less sustainable.<br />

There is nothing new about all this, as described so<br />

eloquently in Peter Matthiessen’s (1975, 1986) eulogies to<br />

the turtle fishermen of Grand Cayman or the striped bass<br />

fishermen of the South Fork of Long Island. However,<br />

societies desperately need to set goals and priorities that<br />

reflect the realities of our lost and dying coral reef resources,<br />

and then follow them. For example, coral reef<br />

fishes may still provide sustainable luxury food and pleasure<br />

for tourists, just as Thompson (1945) proposed, and<br />

captive breeding of aquarium fishes is almost certainly<br />

a viable enterprise. Deciding on such options to the exclusion<br />

of others, and then enforcing them, involves extremely<br />

difficult economic, political and social issues beyond the<br />

bounds of ecological science. But some such decisions are<br />

long overdue, and no more research is required to get<br />

started, because the facts about ‘‘sustainability’’ have been<br />

clear for more than one hundred years.<br />

Getting things straight is all the more important because,<br />

as far as we can tell, almost all the reef species are<br />

still there almost everywhere. The Caribbean monk seal is<br />

extinct and manatees are nearly gone, but even green<br />

turtles still number in the tens of thousands in the Caribbean.<br />

This means that it is still possible, at least in principle,<br />

to save Caribbean coral reefs; although continued<br />

human population growth makes this more and more<br />

unlikely. For example, all the species of corals encountered<br />

previously on Panamanian Caribbean reefs are still<br />

there, even on some of the most devastated reefs, despite<br />

huge decreases in coral abundance (Guzmán and Jackson,<br />

unpublished data). Thus the situation is like the East<br />

African savannas where large animals are more and more<br />

restricted and diminished, but the great majority of species<br />

alive during the Pleistocene still survive (Sinclair and<br />

Arcese 1995).<br />

For this reason, and by analogy to the Serengeti, really<br />

large marine protected areas on the scale of hundreds to<br />

thousands of square kilometers are vital to any hope of<br />

conserving Caribbean coral reefs and coral reef species.<br />

Can we restore damaged reefs? Can we control inputs<br />

from the land and harvesting? Can we manage what we do<br />

decide to invest in and use? These are the questions that<br />

really do merit more research on a monumental scale<br />

(NRC 1995). The people trying to answer them are the<br />

heros of our discipline and the only chance we have got.<br />

Acknowledgements. N. Knowlton talked me into writing all this<br />

depressing stuff down and helped in innumerable other ways. Conversations<br />

with P. Dayton, L. D’Croz, H. Guzma´ n, J. Lang, J.<br />

Ogden, R. Robertson, A. Rodaniche, and J. Wulff helped greatly to<br />

focus my approach. A. Herrera and A. Aguirre tracked down countless<br />

obscure references which the Smithsonian Libraries obtained<br />

with their customary efficiency and grace. F. Herrera kindly provided<br />

data for Kuna populations inside the Comarca which are not<br />

separated from other mainland Kuna populations in the Government<br />

of Panama statistics. N. Knowlton, H. Lessios, D. Levitan and<br />

one other reviewer criticized the manuscript and made many helpful<br />

suggestions. To all I am very grateful.<br />

References<br />

Agassiz A (1883) Explorations of the surface fauna of the Gulf<br />

Stream, under the auspices of the United States Coast Survey. II.<br />

The Tortugas and Florida reefs. Mem Am Acad Arts Sci Centennial<br />

2 : 107—132<br />

Bak RPM (1987) Effects of chronic oil pollution on a Caribbean<br />

coral reef. Mar Poll Bull 18 : 534—539<br />

Beebe W (1928) Beneath tropic seas. Blue Ribbon Books, New York<br />

Bell PRF, Tomascik T (1993) The demise of the fringing coral reefs<br />

of Barbados and of regions in the Great Barrier Reef (GBR)<br />

lagoon-impacts of eutrophication. In: Ginsburg RN (comp) Proceedings<br />

of the colloquium on global aspects of coral reefs:<br />

health, hazards and history. Rosenstiel School Mar Atmos Sci,<br />

University of Miami, Florida, pp 319—325<br />

Bernáldez A (1988) History of the Catholic sovereigns, Don Ferdinand<br />

and Don a Isabella, chapters 123—131. In: Jane C (ed) The<br />

four voyages of Columbus. Dover Publications, New York,<br />

pp 115—188<br />

Bjorndal KA (1982) The consequences of herbivory for the life<br />

history pattern of the Caribbean green turtle, Chelonia mydas. In:<br />

Bjorndal KA (ed) Biology and conservation of sea turtles. Smithsonian<br />

Institution Press, Washington, DC, pp 111—116<br />

Bjorndal KA, Zug GR (1995) Growth and age of sea turtles (revised<br />

edn). In: Bjorndal KA (ed) Biology and conservation of<br />

sea turtles. Smithsonian Institution Press, Washington, DC,<br />

pp 599—600


Bradbury RH, Seymour RM, Antonelli PL (1992) Is the Great<br />

Barrier Reef ecologically sustainable? Proc 7th Int Coral Reef<br />

Symp 2 : 803<br />

Carr AF (1956) The windward road, Alfred Knopf, New York<br />

Chapin M (1995) Epilogue. In: Roeder H, (ed) Plants and animals<br />

in the life of the Kuna. University of Texas Press, Austin,<br />

pp 115—119<br />

Clark HL (1919) The distribution of the littoral echinoderms of the<br />

West Indies. Pap Dept Mar Biol Carnegie Inst Wash 13 : 51—73<br />

Clifton KE, Kim K, Wulff JL (in press) A field guide to the reefs of<br />

Caribbean Panama with an emphasis on western San Blas. Proc<br />

8th Int Coral Reef Symp<br />

Connell JH (1978) Diversity in tropical rain forests and coral reefs.<br />

Science 199 : 1302—1310<br />

Cortés J (1993) A reef under siltation stress: a decade of degradation.<br />

In: Ginsburg RN (comp) Proceedings of the colloquium on<br />

global aspects of coral reefs: health, hazards and history. Rosentiel<br />

School Mar Atmos Sci, University of Miami, Florida,<br />

pp 240—246<br />

Cortés J, Risk MJ (1985) A reef under siltation stress: Cahuita, Costa<br />

Rica. Bull Mar Sci 36 : 339—356<br />

Dampier W (<strong>17</strong>29) A new voyage around the world (1968 reprint).<br />

Dover, New York<br />

Dodge RE, Vaisnys JR (1977) Coral populations and growth patterns:<br />

responses to sedimentation and turbidity associated with<br />

dredging. J Mar Res 35 : 715—730<br />

Donovan SK, Gordon CM (1993) Echinoid taphonomy and the<br />

fossil record: supporting evidence from the Plio-Pleistocene of<br />

the Caribbean. Palaios 8 : 304—306<br />

Dublin HT (1995) Vegetation dynamics in the Serengeti-Mara ecosystem:<br />

the role of elephants, fire, and other factors. In: Sinclair<br />

ARE, Arcese P (eds) Serengeti II: dynamics, management, and<br />

conservation of an ecosystem. University of Chicago Press,<br />

Chicago, pp 71—90<br />

Duerden JE (1901) The marine resources of the British West Indies.<br />

West Indian Bull: J Imperial Agric Dept West Indies<br />

1901 : 121—141<br />

Elton C (1927) Animal ecology. Macmillan, New York<br />

Field GW (1891) Notes on the echinoderms of Kingston Harbor,<br />

Jamaica, W. I. The Johns Hopkins Univ Circ 2 : 83—84<br />

Gardner WJ (1909) A history of Jamaica. T. Fisher Unwin, London<br />

Glynn PW (1973) Aspects of the ecology of reefs in the western<br />

Atlantic region. In: Jones OA, Endean R (eds) Biology and<br />

geology of coral reefs, vol 2. Academic Press, New York,<br />

pp 271—324<br />

Gordon CM, Donovan SK (1992) Disarticulated echinoid ossicles in<br />

paleoecology and taphonomy: the last interglacial Falmouth<br />

Formation of Jamaica. Palaios 7 : 157—166<br />

Goreau TF (1959) The ecology of Jamaican coral reefs. I. Species<br />

composition and zonation. Ecology 40 : 67—90<br />

Guzmán HM, Weil E, Jackson JBC (1991) Short-term ecological<br />

consequences of a major oil spill on Panamanian subtidal reef<br />

corals. Coral Reefs 10 : 1—12<br />

Guzmán HM, Burns KA, Jackson JBC (1994) Injury, regeneration<br />

and growth of Caribbean reef corals after a major oil spill in<br />

Panama. Mar Ecol Prog Ser 105 : 231—241<br />

Hay ME (1984) Patterns of fish and urchin grazing on Caribbean<br />

coral reefs: are previous results typical? Ecology 65 : 446—454<br />

Hay ME (1991) Marine-terrestrial contrasts in the ecology of plant<br />

chemical defenses against herbivores. Trends Ecol Evol<br />

6 : 362—365<br />

Hay ME (in press) Seaweeds and the ecology and evolution of coral<br />

reefs. Coral Reefs<br />

Heinsohn GE, Wake J, Marsh H, Spain AV (1977) The dugong<br />

(Dugong dugon (Muller)). Biol Conserv 6 : 143—152<br />

Henderson JB (1914) The cruise of the Tomas Bárrera. GP Putnam’s<br />

Sons, New York<br />

Hughes TP (1989) Community structure and diversity of coral reefs:<br />

the role of history. Ecology 70 : 275—279<br />

Hughes TP (1994) Catastrophes, phase shifts, and large-scale degradation<br />

of a Caribbean coral reef. Science 265 : 1547—1551<br />

S31<br />

Hutchinson GE (1978) An introduction to population ecology. Yale<br />

University Press, New Haven, Connecticut<br />

Jackson JBC (1982) Biological determinants of present and past<br />

sessile animal distributions. In: Tevesz MJS, McCall PL (eds)<br />

Biotic interactions in recent and fossil benthic communities.<br />

Plenum, New York, pp 39—120<br />

Jackson JBC (1991) Adaptation and diversity of reef corals. Bio-<br />

Science 41 : 475—482<br />

Jackson JBC (1992) Pleistocene perspectives on coral reef community<br />

structure. Am Zool 32 : 719—731<br />

Jackson JBC (1995) The role of science in coral reef conservation<br />

and management. In: Partnership building and framework development.<br />

Final report, the <strong>International</strong> Coral Reef Initiative<br />

Workshop, Dumaguete City, Philippines, pp 5—9<br />

Jackson JBC, Cubit JD, Keller BD, Batista V, Burns K, Caffey HM,<br />

Caldwell RL, Garrity SD, Getter CD, Gonzalez C, Guzmán HM,<br />

Kaufmann KW, Knap AH, Levings SC, Marshall MJ, Steger R,<br />

Thompson RC, Weil E (1989) Ecological effects of a major oil<br />

spill on Panamanian coastal marine communities. Science<br />

243 : 37—44<br />

Jackson JBC, Budd AF, Pandolfi JM (1996) The shifting balance of<br />

natural communities? In: Jablouski D, Erwin DH, Lipps, JH<br />

(eds) Evolutionary paleobiology, Universiy of Chicago Press,<br />

Chicago, pp 89—122<br />

Janzen DH, Martin PS (1982) Neotropical anachronisms: the fruits<br />

the gomphotheres ate. Science 215 : 19—27<br />

Janzen DH, Wilson DE (1983) Mammals: introduction. In: Janzen<br />

DH (ed) Costa Rican natural history, University of Chicago<br />

Press, Chicago, pp 426—442<br />

King FW (1982) Historical review of the decline of the green turtle<br />

and the hawksbill. In: Bjorndahl KA (ed) Biology and conservation<br />

of sea turtles. Smithsonian Institution Press, Washington,<br />

DC, pp 183—188<br />

Knowlton N (1992) Thresholds and multiple stable states in coral<br />

reef community dynamics. Am Zool 32 : 674—682<br />

Knowlton N, Lang JC, Keller BD (1990) Case study of natural<br />

population collapse: post-hurricane predation on Jamaican<br />

staghorn corals. Smithson Contrib Mar Sci 31 : 1—25<br />

Lanyon JM, Limpus CJ, Marsh H (1989) Dugongs and turtles:<br />

grazers in the seagrass system. In: Larkum AWD, McComb AJ,<br />

Sheperd SA (eds) Biology of seagrasses: a treatise on the biology<br />

of seagrasses with special reference to the Australian region,<br />

Elsevier, Amsterdam, pp 610—634<br />

Lessios HA (1988) Mass mortality of Diadema antillarum in the<br />

Caribbean: what have we learned? Ann Rev Ecol Syst 19 : 371—393<br />

Lessios HA, Robertson DR, Cubit JD (1984) Spread of Diadema<br />

mass mortality through the Caribbean. Science 226 : 335—337<br />

Levitan DR (1991) Skeletal changes in the test and jaws of the sea<br />

urchin Diadema antillarum in response to food limitation. Mar<br />

Biol 111 : 431—435<br />

Levitan DR (1992) Community structure in times past: influence of<br />

human fishing pressure on algal-urchin interactions. Ecology<br />

73 : 1597—1605<br />

Lewis BC (1940) The Cayman Islands and marine turtle. Bull Inst<br />

Jamaica, Sci Ser 2 : 56—65<br />

Lewis JB (1960) The coral reefs and coral communities of Barbados,<br />

W. I. Can J Zool 38 : 1133—1145<br />

Lewis JB (1984) The Acropora inheritance: a reinterpretation of the<br />

development of fringing reefs in Barbados, West Indies. Coral<br />

Reefs 3 : 1<strong>17</strong>—122<br />

Lewsey CD (1978) Assessing the environmental effects of tourism on<br />

the carrying capacity of small island systems ‘‘The case for<br />

Barbados’’. ‘‘Unpubl. Ph.D Diss, Cornell University, Univ<br />

Microfilms Int, Ann Arbor, Michigan<br />

Liddell WD, Ohlhorst SL (1988) Comparison of western Atlantic<br />

coral reef communities. Proc 6th Int Coral Reef Symp 3 : 281—286<br />

Limpus CJ (1995) Global overview of the status of marine turtles:<br />

a 1995 viewpoint. In: Bjorndal KA (ed), Biology and conservation<br />

of sea turtles, 2nd edn. Smithsonian Institution Press, Washington<br />

DC, pp 605—609


S32<br />

Long E (<strong>17</strong>74) The history of Jamaica, or general survey of the<br />

antient and modern state of that island: with reflections on its<br />

situations, settlements, inhabitants, climate, products, commerce,<br />

laws and government (1970 Reprint). Frank Cass and Co,<br />

London<br />

Lugo A, Schmidt R, Brown S (1981) Tropical forests in the Caribbean.<br />

Ambio 10 : 318—324<br />

Macintyre IG (1968) Preliminary mapping of the insular shelf off the<br />

west coast of Barbados. Carib J Sci 8 : 95—100<br />

Matthiessen P (1975) Far Tortuga. Random House, New York<br />

Matthiessen P (1986) Men’s lives: the surfmen and the baymen of the<br />

South Fork. Random House, New York<br />

McClanahan TR, Muthiga NA (1988) Changes in Kenyan coral reef<br />

community structure and function due to exploitation.<br />

Hydrobiologia 166 : 269—276<br />

McClanahan TR, Kamukuru AT, Muthiga NA, Gilagabher Yebio<br />

M, Obura D (1996) Effect of sea urchin reductions on algae,<br />

coral, and fish populations. Conserv Biol 10 : 136—154<br />

McNaughton SJ, Banyikwa FF (1995) Plant communities and herbivory.<br />

In: Sinclair ARE, Arcese P (eds) Serengeti II: dynamics,<br />

management, and conservation of an ecosystem. University of<br />

Chicago Press, Chicago, pp 49—70<br />

Mesolella KJ (1957) Zonation of uplifted Pleistocene coral reefs on<br />

Barbados, West Indies. Science 156 : 638—640<br />

Meylan A (1985) Nutritional characteristics of sponges in the diet of<br />

the hawksbill turtle, Eretmochelys imbricata 3rd Int Sponge Conf:<br />

472—477<br />

Meylan A (1988) Spongivory in hawksbill turtles: a diet of glass.<br />

Science 239 : 393—395<br />

Munroe JL (ed) (1983) Caribbean coral reef fisheries, 2nd ed.<br />

ICLARM Stud Rev 7 : 1—276<br />

National Research Council (1995) Understanding marine biological<br />

diversity. National Academy Press, Washington, DC<br />

National Geographic Society (1981) Geographic atlas of the world,<br />

5th edn. Nat Geogr Soc, Washington, DC<br />

Neitschman B (1973) Between land and water: the subsistence ecology<br />

of the Moskito Indians, eastern Nicaragua. Seminar Press,<br />

New York<br />

Neitschman B (1982) The cultural context of sea turtle subsistence<br />

hunting in the Caribbean and problems caused by commercial<br />

exploitation. In: Bjorndal KA (ed), Biology and conservation<br />

of sea turtles. Smithsonian Institution Press, Washington, DC,<br />

pp 439—445<br />

Nutting CC (1919) Barbados-Antigua expedition. Univ Iowa Studies<br />

Nat Hist 8 : 1—274<br />

Ogden JC (1980) Faunal relationships in Caribbean seagrass beds.<br />

In: Phillips RC, McRoy CP (eds), Handbook of seagrass biology:<br />

an ecosystem perspective, Garland STPM Press, New York,<br />

pp <strong>17</strong>3—198<br />

Ogden JC, Robinson L, Whitlock H, Daganhart H, Cebula R (1983)<br />

Diel foraging patterns in juvenile green turtles (Chelonia mydas<br />

L) in St. Croix, United States, Virgin Islands. J Exp Mar Biol<br />

Ecol 66 : 199—205<br />

Ogden JC, Ogden NB (1993) The coral reefs of the San Blas Islands:<br />

revisited after 20 years. In: Ginsburg RN (comp) Proceedings of<br />

the colloquium on global aspects of coral reefs: health, hazards,<br />

and history. Rosentiel School Mar Atmos Sci, University of<br />

Miami, Florida, pp 267—272<br />

Pauly D (1995) Anecdotes and the shifting baseline syndrome of<br />

fisheries. Trends Ecol Evol 10 : 430<br />

Porter JW (1972) Ecology and species diversity of coral reefs on<br />

opposite sides of the Isthmus of Panamá. Bull Biol Soc Wash<br />

2:89—116<br />

Randall JE (1965) Grazing effect on sea grasses by herbivorous reef<br />

fishes in the West Indies. Ecology 46 : 255—260<br />

Randall JE, Schroeder RE, Starck WA, Jr (1961) Notes on the<br />

biology of Diadema antillarum. Carib J Sci 4 : 421—433<br />

Randall JE, Hartman WD (1968) Sponge-feeding fishes of the West<br />

Indies. Mar Biol 1 : 216—225<br />

Roberts L (1989) Disease and death in the New World. Science<br />

246 : 1245—1247<br />

Rodriguez A (1981) Marine and coastal environmental stress in the<br />

wider Caribbean region. Ambio 10 : 283—294<br />

Rogers CS (1985) Degradation of Caribbean and western Atlantic<br />

coral reefs and decline of associated fisheries. Proc 5th Int Coral<br />

Reef Congr 6 : 491—496<br />

Rogers CS (1990) Responses of coral reefs and reef organisms to<br />

sedimentation. Mar Ecol Prog Ser 62 : 185—202<br />

Rouse I (1992) The Tainos. Yale University Press, New Haven<br />

Sauer CO (1966) The early Spanish Main. University of California<br />

Press, Berkeley, California<br />

Sheppard C (1995) The shifting baseline syndrome. Mar Poll Bull<br />

30 : 766—767<br />

Sheppard C, Price A, Roberts C (1992) Marine ecology of the<br />

Arabian region. Academic Press, London<br />

Sinclair ARE (1995a) Serengeti past and present. In: Sinclair ARE,<br />

Arcese P (eds) Serengeti II: dynamics, management, and conservation<br />

of an ecosystem. University of Chicago Press, Chicago,<br />

pp 3—30<br />

Sinclair ARE (1995b) Equilibria in plant-herbivore interactions. In:<br />

Sinclair ARE, Arcese P (eds) Serengeti II: dynamics, management,<br />

and conservation of an ecosystem. University of Chicago<br />

Press, Chicago, pp 91—113<br />

Sinclair ARE, Arcese P (1995) Serengeti II: dynamics, management,<br />

and conservation of an ecosystem. University of Chicago Press,<br />

Chicago<br />

Sloane H (<strong>17</strong>07—<strong>17</strong>25) A voyage to the islands Madera, Barbadoes,<br />

Nieves, St Christophers, and Jamaica; with the natural history of<br />

the herbs and trees, four-footed beasts, fishes, birds, insects,<br />

reptiles, &c. of the last of those islands. In two volumes. Printed<br />

for the author, London<br />

Stearn CW, Scoffin TP, Martindale W (1977) Calcium, carbonate<br />

budget of a fringing reef on the west coast of Barbados.<br />

I. Zonation and productivity. Bull Mar Sci 27 : 479—510<br />

Thayer GW, Engel DW, Bjorndahl KA (1982) Evidence for shortcircuiting<br />

of the detritus cycle of seagrass beds by the green turtle,<br />

Chelonia mydas. J Exp Mar Biol Ecol 62 : <strong>17</strong>3—183<br />

Thayer GW, Bjorndahl KA, Ogden JC, Williams SL, Zieman JC<br />

(1984) Role of larger herbivores in seagrass communities. Estuaries<br />

7 : 351—376<br />

Thompson EF (1945) The fisheries of Jamaica. Development and,<br />

welfare in the West Indies, Bulletin 18, Bridgetown, Barbados<br />

Tomascik T, Sander F (1985) Effects of eutrophication on reefbuilding<br />

corals. I. Growth rate of the reef building coral Montastrea<br />

annularis. Mar Biol 87 : 143—155<br />

Tomascik T, Sander F (1987a) Effects of eutrophication on reefbuilding<br />

corals. II. Structure of scleractinian coral communities<br />

on fringing reefs, Barbados, West Indies. Mar Biol 944 : 3—75<br />

Tomascik T, Sander F (1987b) Effects of eutrophication on reefbuilding<br />

corals. III. Reproduction of the reef-building coral<br />

Porites porites. Mar Biol 94 : 77—94<br />

van Dam RP, Diez CE (in press) Predation by hawksbill turtles on<br />

sponges at Mona Island, Puerto Rico. Proc 8th Int Coral Reef<br />

Symp<br />

Ventocilla J, Herrera H, Nu´ nez V (1995) Plants and animals in the<br />

life of the Kuna, Roeder H (ed) University of Texas Press, Austin,<br />

150 pp<br />

Wilkinson CR (1992) Coral reefs of the world are facing widespread<br />

devastation: can we prevent this through sustainable management<br />

practices. Proc 7th Int Coral Reef Symp 1 : 11—21<br />

Woodley JD, Chornesky EA, Clifford PA, Jackson JBC, Kaufman<br />

LS, Knowlton N, Lang JC, Pearson MP, Porter JW, Rooney<br />

MC, Rylaarsdam KW, Tunnicliffe VJ, Wahle CM, Wulff JL,<br />

Curtis ASG, Dallmeyer MD, Jupp BJ, Koehl MAR, Neigel J,<br />

Sides EM (1981) Hurricane Allen’s impact on Jamaican coral<br />

reefs. Science 214 : 749—755<br />

Young T (1847) Narrative of a residence on the Mosquito shore:<br />

with an account of Truxillo, and the adjacent islands of Bonacca<br />

and Roatan; and a vocabulary of the Mosquitian language.<br />

Smith, Elder, London (1971 reprint). Krause Reprint, New York<br />

Ziemann (1982) The ecology of seagrasses of south Florida: a community<br />

profile. US Fish Wildlife Serv Prog FWS/OBS-82/25 124 : 1—26


Social Interactions in Captive Female<br />

Florida Manatees<br />

Jennifer Young Harper and Bruce A. Schulte n<br />

Zoo Biology 24:135<strong>–</strong>144 (2005)<br />

Department of Biology, Georgia Southern University, Statesboro, Georgia<br />

The Florida manatee (Trichechus manatus latirostris) is considered a semi-social<br />

species with strong bonds developed primarily between mother and offspring.<br />

Some field studies suggest sociality may be more developed and such social<br />

relationships may facilitate survival. Seven facilities in Florida house manatees,<br />

many of which were brought into captivity because of injury or illness sustained in<br />

the wild. Decisions to release such manatees consider individual history and<br />

health. We examined social interactions in adult female captive manatees to assess<br />

level of association and implications for manatee care and release. We<br />

investigated the degree of contact among 20 manatees in captivity at four<br />

facilities housing two to nine adult female manatees. We used all contact behavior<br />

occurrences sampling and continuous recording for 180 continuous minutes per<br />

day over 3 consecutive days at each facility. Virtually all contacts were nonaggressive.<br />

The number of contacts between manatees increased as the number of<br />

manatees per unit volume of water increased. Contacts did not fit a Poisson<br />

distribution, however, and were not random. When more than two manatees were<br />

present, manatees only associated with a subset of individuals in the aquarium.<br />

Relationships maintained in captivity indicate the potentially social nature of<br />

manatees, and suggest that further research is needed to examine the benefit of<br />

these relationships to the health and rehabilitation of manatees in captivity and<br />

conservation in the wild. Zoo Biol 24:135<strong>–</strong>144, 2005. c 2005 Wiley-Liss, Inc.<br />

Key words: aggression; association; conservation; ethogram; interaction; social behavior<br />

INTRODUCTION<br />

The organization of individuals into social groups is likely to evolve when the<br />

benefits of group living outweigh the costs [Alexander, 1974; Emlen 1997]. Benefits of<br />

Dr. Jennifer Young Harper’s present address is Coastal Georgia Community College, 3700 Altama<br />

Avenue, Brunswick, GA 31520.<br />

n<br />

Correspondence to: Bruce A. Schulte, Department of Biology PO Box 8042, Georgia Southern<br />

University, Statesboro, Georgia, 30460-8042. E-mail: bschulte@georgiasouthern.edu<br />

Received 2 April 2004; Accepted 8 November 2004<br />

DOI 10.1002/zoo.20044<br />

Published online in Wiley InterScience (www.interscience.wiley.com).<br />

c 2005 Wiley-Liss, Inc.


136 Harper and Schulte<br />

group living include greater access to resources and protection. Group living has<br />

ecological costs including an increase in competition for food, space, and mates and<br />

exposure to parasites, pathogens, and predators. Conflicts and aggression can<br />

become heightened in crowded areas, especially when there is competition over a<br />

limited resource. The resolution of conflict can occur when individuals leave a<br />

particular situation, defend territories, form small closely associated groups, or<br />

establish dominance hierarchies built upon aggression or affiliation [Maher and<br />

Lott, 1995, 2000; Pusey and Packer, 1999].<br />

Florida manatees (Trichechus manatus latirostris), a subspecies of the West<br />

Indian manatee, are considered semi-social or asocial animals, because they<br />

generally do not form organized social structures. Manatees are usually seen<br />

traveling in warmer coastal waters alone or in mother<strong>–</strong>calf pairs, with such bonds<br />

lasting until the calf is approximately 1<strong>–</strong>2 years old [Reynolds and Odell, 1991].<br />

From 1993<strong>–</strong>1996, Koelsch [1997] observed manatees in Sarasota Bay, Florida and<br />

noted some strong bonds among adult manatees. In addition, associations were<br />

stronger within than between the sexes. These non-random association patterns<br />

along with the social facilitation noted by Reynolds [1981] suggest that manatees<br />

may be more social than previously thought [Koelsch, 1997]. The current study<br />

investigated the type and extent of interactions among female manatees held in<br />

captivity to evaluate the strength of social bonds.<br />

Manatees in the wild face many problems in their aquatic environment<br />

including ingestion of foreign materials [Beck and Barros, 1991], exposure to toxins<br />

such as red tide [Bossart et al., 1998], and collisions with boats [Wright et al., 1995].<br />

Injured or sick manatees are taken into captivity for rehabilitation. Once an injured<br />

animal is out of critical care in a captive facility, it is usually housed with several<br />

other manatees of the same sex. We know relatively little, however, about how these<br />

supposed semi-social animals behave in the constant presence of multiple<br />

conspecifics. The specific objective of this research was to determine if manatees in<br />

captivity exhibit behaviors that are either aggressive, serve to reduce aggression or<br />

reflect the establishment of particular associations. Manatees were observed before,<br />

during, and after feeding periods because conflicts might arise in anticipation of, or<br />

during consumption of, food (a potentially limited resource). We examined two<br />

hypotheses on level of interaction: (1) manatees will rarely interact; and (2) manatees<br />

will interact more under higher density situations (i.e., more manatees per unit<br />

volume of aquarium space). In addition, we evaluated two hypotheses on the timing<br />

and type of interactions. Specifically, (3) most interactions will be aggressive and<br />

occur just before feeding or when food is becoming depleted, and (4) if manatees are<br />

more social than previously thought, most interactions will be non-aggressive and<br />

occur outside of feeding periods.<br />

MATERIALS AND METHODS<br />

Study Sites<br />

This research was conducted from January<strong>–</strong>March 2001 at four zoos and<br />

aquariums housing captive manatees, specifically Homosassa Springs Wildlife State<br />

Park (HSWSP, Homosassa, Florida), Lowry Park Zoo (Tampa, Florida), Sea World<br />

of Florida (Orlando, Florida), and Walt Disney World’s The Living Seas in Epcot


Captive Female Manatee Social Behavior 137<br />

(Orlando, Florida). Generally, manatees were housed in human-made enclosures<br />

approximately 3 m in depth and oval in shape. The exception was HSWSP, which is<br />

a naturally occurring spring that covers approximately 0.2 ha and reaches depths of<br />

13.5 m. Manatees (n ¼ 20) at all facilities were viewed from an underwater<br />

observation area or from above the enclosure. We estimated the volume of water<br />

for HSWSP from the dimensions of the exhibit. The other three facilities provided us<br />

with aquarium volumes. Water volumes at each facility are as follows: HSWSP (1.1<br />

10 7 L [9 manatees]), Sea World (1.4 10 6 L [6 manatees]), Epcot (760,000 L<br />

[2 manatees]), and Lowry Park Zoo (470,000 L [3 manatees]).<br />

Behavioral Observations<br />

Manatees were watched continuously for 180 min over 3 days around<br />

feeding times. The time duration was selected because there were constraints<br />

at the facilities that prevented longer observation periods including operating<br />

hours, interference in observations when visitor numbers were high, and<br />

logistical issues. Observations were made at the four facilities with 20 individuals<br />

for a total of 36 hr. No adult males were observed for this study, but three<br />

juvenile male manatees were observed. One male was at Lowry Park Zoo<br />

and two were at Sea World. All three males were with females. During the<br />

3 hr of continuous observations for all interactions, we recorded all aggressive<br />

and non-aggressive contacts between manatees (sender<strong>–</strong>receiver) and spatial<br />

displacement by manatees [Martin and Bateson, 1993]. Aggressive interactions were<br />

defined as contacts in which the sender’s head touched the receiver with some<br />

momentum or caused physical movement of the receiver. Displacement was occurred<br />

when one manatee (sender) caused another manatee (receiver) to move from its<br />

location without contact because of the presence or movement of the sender.<br />

Displacements were rare and not considered in the analysis. Other aggressive<br />

interactions occurred when the sender’s tail hit the receiver with force as it swam<br />

away. Non-aggressive interactions that might be affiliative in nature included all<br />

other contacts (head to different body parts, body to different body parts, fin to<br />

different body parts, and tail to different body parts without force). Incidental<br />

contacts transpired when an animal brushed against another animal. These contacts<br />

were not included in the analysis.<br />

Manatees at all facilities were fed 25<strong>–</strong>100 heads of romaine lettuce<br />

depending on the number of manatees in the aquarium (one head weighs<br />

approximately 488 grams; V. Burke, head keeper at Lowry Park Zoo,<br />

personal communication). Feeding occurred three to four times a day at regular<br />

intervals. Three observational periods were defined as pre-feeding, during-feeding,<br />

and post-feeding. The pre-feeding period occurred during the first 60 min of<br />

observations before the handlers provisioned any food. The feeding period occurred<br />

when the manatees were feeding and post-feeding was defined as when 10% or less of<br />

the food remained (heads and leaves combined). The number of heads of lettuce<br />

(or leaf equivalents) that remained in the aquarium were counted and recorded every<br />

5 min. Feeding periods lasted approximately 1 hr. Supplemental foods such as<br />

carrots, kale, and apples were sometimes provided to manatees by the aquarium<br />

staff, but not included in the determination of feeding or non-feeding observational<br />

periods.


138 Harper and Schulte<br />

Data Analysis<br />

The total number of contacts for each day and facility was counted. The<br />

proportion of aggressive or non-aggressive interactions by manatees at each facility<br />

was calculated by observational period over the 3 consecutive days of observations.<br />

Similar proportions also were calculated for individual manatees as sender and as<br />

receiver. Sender-receiver matrices were created for each facility. Non-aggressive<br />

contacts were used to construct sender-receiver matrices because of the low<br />

occurrence of aggressive contacts. Animals were ordered in the matrices based on the<br />

total number of contacts sent. Log-linear G-test analyses were used to examine if<br />

non-aggressive contacts were more likely before or after feeding at separate facilities<br />

and to compare the distribution among facilities of non-aggressive contacts before<br />

and after feeding. Linear regression was used to determine the effect of density on<br />

total number of contacts. We calculated the coefficient of dispersion (CD) using the<br />

number of contacts between unique pairs of manatees to determine if contacts fit a<br />

Poisson distribution, where values greater than one show a clumped distribution,<br />

whereas values less than one indicate repulsion. For the two facilities (HSWSP and<br />

SeaWorld) with greater than three manatees, we analyzed the distribution of these<br />

contacts using a G-test with the Williams correction factor [Sokal and Rohlf, 1995].<br />

The distribution is determined by the frequency of manatee pairs with zero to the<br />

maximum number of observed contacts. To maintain frequencies of at least four<br />

occurrences per cell (i.e., contacts/manatee pair) some categories were combined. In<br />

the two analyses, we used three categories. Because the parameter mean in each case<br />

was estimated from the sample, two degrees of freedom were removed, yielding a<br />

single degree of freedom for the G-statistic [Sokal and Rohlf, 1995]. Significant<br />

variation from a Poisson distribution indicated a non-random pattern of contacts.<br />

Analyses were carried out using JMP, version 3.02 (SAS Institute Inc., Cary, NC),<br />

and by methods described by Sokal and Rohlf [1995].<br />

RESULTS<br />

Manatees interacted regularly during the non-feeding periods of our study. We<br />

observed no contacts between manatees while they were feeding. Of the 218 nonaggressive<br />

encounters, 98 (45%) occurred in the pre-feeding period and 120 (55%) in<br />

the post-feeding period. The pre-feeding period was always 60 min long whereas the<br />

post period ranged from 30<strong>–</strong>75 min (55.773.62 min, mean7SE). The number of<br />

non-aggressive contacts did not differ in the pre- vs. the post-feeding observational<br />

period within each facility (all Go1.1, df ¼ 1, P40.05). The distribution of nonaggressive<br />

contacts also did not differ in the pre- and post-feeding periods among the<br />

four facilities (G ¼ 2.02, df ¼ 3, P40.05). The number of total contacts per day<br />

ranged from 10<strong>–</strong>34 at each of the four facilities (means7SE of contacts/hr: Day<br />

1 ¼ 6.7572.11, Day 2 ¼ 6.371.48, Day 3 ¼ 5.970.89).<br />

Of the 228 contacts recorded, only 10 (4.39%) were aggressive. The manatees<br />

at Lowry Park Zoo accounted for nine of 10 aggressive encounters and seven were<br />

observed during a single 180-min observation session. These contacts were head-tobody<br />

contacts mostly by the adult females to a 9-month-old orphaned male. Using<br />

sender-receiver matrices based upon non-aggressive contacts, we discuss manatee<br />

interactions at each facility.


Captive Female Manatee Social Behavior 139<br />

Epcot: The Living Seas<br />

Walt Disney World’s The Living Seas in Epcot housed two female manatees,<br />

Lydia and Mariah. Forty-three non-aggressive contacts (4.871.3 contacts/hr)<br />

occurred between these two manatees. Lydia initiated 20 contacts and Mariah<br />

initiated 23.<br />

Homosassa Springs Wildlife State Park<br />

Homosassa Springs Wildlife State Park housed nine female manatees:<br />

Amanda, Ariel, Betsy, Electra, Holly, Lorelei, Oakley, Rosie, and Willoughby.<br />

Thirty-one non-aggressive contacts (3.470.1 contacts/hr) occurred between the<br />

manatees. Eighteen contacts were made between three related individuals (Amanda,<br />

Ariel and Betsy) termed Group one, consisting of a mother and her two adult<br />

daughters. The remaining six manatees made 12 contacts to each other with only one<br />

contact made between the two groups of manatees (Table 1).<br />

Lowry Park Zoo<br />

Lowry Park Zoo housed two female manatees, Cinco and BB, and one<br />

orphaned male calf, Lowry. Eighty-three contacts occurred between these manatees;<br />

nine were aggressive (10.8%, 1.070.67 contacts/hr) and 74 were non-aggressive<br />

(89.2%, 8.270.9 contacts/hr). BB initiated five aggressive contacts to Lowry. Both<br />

Cinco and Lowry made two aggressive contacts (22.2%) toward each other. Lowry<br />

initiated the most non-aggressive contacts (38). BB and Cinco initiated 20 and 16<br />

non-aggressive contacts, respectively. For the adult females, BB initiated contact<br />

with Cinco eight times and Cinco contacted BB nine times.<br />

Sea World<br />

Sea World of Florida housed four adult female manatees: Charlotte, Rita,<br />

Sara, and Stubbie and two orphaned male calves: Brooks and Pistachio. Of the 71<br />

interactions that occurred, 70 (98.6%, 7.871.0 contacts/hr) were non-aggressive and<br />

one (1.4%, 0.<strong>17</strong>0.1 contact/hr) was aggressive. The juvenile males initiated more<br />

contacts than the adult females (52 of 71 or 73%). Brooks made 34 contacts with<br />

most directed toward Charlotte (15) and Sara (13). These were primarily play<br />

mounting contacts with the two females. Pistachio initiated four similar contacts<br />

TABLE 1. Matrix constructed from non-aggressive contacts at Homosassa Springs Wildlife<br />

State Park with nine adult females<br />

Amanda a Rosie Betsy a Ariel a Holly Will Oakley Lorelei Electra Total<br />

Amanda — 0 2 7 0 0 0 1 0 10<br />

Rosie 0 — 0 0 1 1 2 0 3 7<br />

Betsy 3 0 — 2 0 0 0 0 0 5<br />

Ariel 2 0 2 — 0 0 0 0 0 4<br />

Holly 0 0 0 0 — 1 0 0 1 2<br />

Will 0 0 0 0 0 — 1 0 0 1<br />

Oakley 0 0 0 0 0 1 — 0 0 1<br />

Lorelei 0 0 0 0 0 1 0 — 0 1<br />

Electra 0 0 0 0 0 0 0 0 — 0<br />

a Amanda, Betsy and Ariel are related as mother and two adult daughters.


140 Harper and Schulte<br />

each to Sara, Rita and Stubbie. Excluding the juvenile males from total contacts,<br />

Charlotte and Sara accounted for 11 of the 13 encounters (Table 2).<br />

All Facilities<br />

All manatees, with the exception of one, were involved in some type of<br />

interaction with other individuals. Density of manatees per unit volume explained<br />

the majority of variation in contact rates (Fig. 1; r 2 ¼ 0.96). The coefficient of<br />

TABLE 2. Matrix constructed from non-aggressive manatee contacts at Sea World of Florida<br />

Brooks a<br />

Pistachio b<br />

Charlotte c<br />

Sara c<br />

Rita c<br />

Stubbie c<br />

Total<br />

Brooks — 3 15 13 2 2 35<br />

Pistachio 3 — 1 4 4 4 16<br />

Charlotte 1 0 — 8 0 0 9<br />

Sara 2 0 3 — 0 0 5<br />

Rita 0 0 1 1 — 0 2<br />

Stubbie 1 0 0 0 0 — 1<br />

a Juvenile manatee (2 years old).<br />

b Juvenile manatee (1.5 years old).<br />

c Adult female manatee.<br />

Total Number of Contacts<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

HSSP<br />

Epcot<br />

SeaWorld<br />

y = 9.9482x + 21.887<br />

R 2 = 0.96<br />

0 1 2 3 4 5 6<br />

Number of Manatees per 1,000,000 L water<br />

Fig. 1. Number of contacts for each facility (Homosassa Springs Wildlife State Park, Walt<br />

Disney World’s The Living Seas at Epcot, Sea World of Florida, and Lowry Park Zoo)<br />

initiated by manatees per million L water within each aquarium.<br />

LPZ<br />

7


Captive Female Manatee Social Behavior 141<br />

dispersion (CD) at the three facilities with at least three manatees was always greater<br />

than one, indicating clumping of the distribution (HSWSP ¼ 3.9, SeaWorld ¼ 6.2,<br />

Lowry Park Zoo ¼ 1.9). The distribution of contacts was significantly different from<br />

a Poisson distribution for HSWSP (Gadj ¼ 9.24w 2 0.05[1] ¼ 3.84) and for SeaWorld<br />

(Gadj ¼ 13.74w 2 0.05[1] ¼ 3.84). We could not perform the analysis for Lowry Park<br />

Zoo because of the low number of manatees, resulting in single frequencies per cell.<br />

DISCUSSION<br />

Manatees in captivity associated and interacted with certain individuals. The<br />

occurrence of regular, non-aggressive contacts by manatees in captivity and the<br />

relative rarity of aggressive contacts support the contention that manatees may be<br />

more social than previously considered [Koelsch, 1997]. Hence, we reject our first<br />

hypothesis that manatees will rarely interact. The number of non-aggressive contacts<br />

increased with density (Fig. 1), supporting our second hypothesis. Aggression was<br />

rare and only occurred at the facilities with the greatest densities (Lowry Park Zoo<br />

and SeaWorld) but these facilities had one or two juvenile males in with the females.<br />

Of the ten total aggressive encounters observed, nine were recorded at Lowry Park<br />

Zoo. These aggressive contacts were directed mainly toward the 9-month-old male<br />

Lowry by the adult females, possibly to prevent the formation of a mother<strong>–</strong>calf<br />

bond. Lowry was born in captivity but his mother Ionia died from her injuries<br />

during rehabilitation when Lowry was introduced to the two resident adult females,<br />

BB and Cinco. Because aggression was rare and not related to feeding, we reject our<br />

third hypothesis and show support for our fourth hypothesis that most interactions<br />

will be non-aggressive and occur outside of feeding periods. A concurrent study<br />

showed that time in captivity or differences in facilities did not affect manatee<br />

activity patterns [Young, 2001].<br />

The density-dependent nature of non-aggressive interactions suggests that<br />

contacts were random and not selective. Interactions were not equally distributed,<br />

however, among manatees at least at the two facilities with more than three<br />

manatees. At HSWSP, greater association was apparent within two social<br />

subgroups. One subgroup consisted of Amanda and her two adult daughters, Ariel<br />

and Betsy. This observation in captivity fits nicely with field studies, where mother<br />

and offspring form the most apparent long-term bonds during calf dependency in the<br />

wild [Hartman, 1979; Bengtson, 1981; Reynolds, 1981]. All three animals were adults<br />

(Betsy is the youngest, born in captivity in 1990). This long-term close association<br />

may be common in the wild or an artifact of captivity. These three individuals<br />

contacted the other subgroup composed of six unrelated manatees only once. In<br />

contrast, this subgroup exhibited 12 contacts within the group and no contacts to<br />

Amanda, Ariel, or Betsy. Random interactions would predict more contacts between<br />

the two subgroups and the analysis supported a clumped distribution of contacts.<br />

This segregation of manatees is especially interesting at HSWSP, which represented<br />

the site with the lowest density of manatees.<br />

At Sea World of Florida, Charlotte and Sara associated regularly with Rita on<br />

the periphery of this pair. The two young males (Brooks and Pistachio) initiated a<br />

majority of the contacts, mostly as play mounting contacts with the females. Between<br />

the females, the overall distribution of contacts was clumped, not random. At Lowry<br />

Park Zoo, the two adult females, BB and Cinco initiated non-aggressive contacts to


142 Harper and Schulte<br />

each other in nearly equal number, while never initiating aggressive contacts to one<br />

another. This contrasted with the relationship to the orphaned, nine-month-old<br />

male, who received aggressive and non-aggressive contacts. The two adult females at<br />

Epcot showed a similarly reciprocal level of non-aggressive contact initiation as<br />

those at Lowry Park Zoo. Such equal initiation of contacts suggests an absence of<br />

dominance but with only a few animals, statistical analysis is not appropriate.<br />

In wild manatees, affiliative social interactions have been observed between<br />

adult females [Koelsch, 1997]. Mothers and calves associate regularly [Hartman,<br />

1979; Reynolds and Odell, 1991; Koelsch, 1997], but it is uncertain if this<br />

relationship lasts until adulthood as exhibited by the captive manatees in subgroup<br />

one at HSWSP. Elephants are the closest living relative of the Sirenia, and they form<br />

close mother<strong>–</strong>calf associations that last into adulthood for females [Buss and Smith,<br />

1966; Eisenberg et al., 1971; Douglas-Hamilton, 1972; Moss 1976; Dublin, 1983].<br />

The calf learns basic life skills from its mother and social group members [Lee and<br />

Moss, 1999], and young males have even been observed play mounting with their<br />

mothers or other females within their group [Gadgil and Nair, 1983]. Similar play<br />

behavior toward adult female manatees was observed with juvenile males at Sea<br />

World. In the wild, juveniles seem to be submissive to adult manatees [Hartman,<br />

1979], perhaps explaining the prevalence of aggressive contacts initiated by the adult<br />

females toward the juvenile male, Lowry. The types of associations observed with<br />

captive manatees are similar to those described by other mammals in captivity<br />

including elephants [Schulte, 2000], zebras [Schilder, 1992], captive spotted hyenas<br />

[Glickman et al., 1997], and bonobos [de Waal, 1995].<br />

In the case of manatees, understanding relationships in captivity and the wild<br />

becomes especially important in the rehabilitation and eventual release of manatees.<br />

Releasing manatees near where they were captured may be beneficial because<br />

tradition is probably important in manatee survival [Bengtson, 1981]. Manatees may<br />

learn feeding and resting areas among other information from older conspecifics<br />

during development. Like elephants [McComb et al., 2001] and cetaceans [Wells<br />

et al., 1999], manatees may have population level, behavioral diversity related to<br />

cultural traditions. Associations between individuals beyond weaning may be vital in<br />

manatee society and enhance survivorship. Such information could be helpful for<br />

rehabilitation and conservation. To our knowledge, in the instances of simultaneous<br />

releases of manatees from captivity, associations established in captivity have not<br />

been maintained (at least five: female and a surrogate calf, two pairs of males, one<br />

female pair and one mixed sex pair; R. Bonde, Sirenia Project, personal<br />

communication). The sample size, however, is relatively small. Our data suggest<br />

that females do establish some particular association groups in captivity, but further<br />

study is needed to assess to what degree such associations are a result of captivity. In<br />

terms of captive manatee management, density did not seem to have any negative<br />

effects on manatee behavior and social aggregation of manatees seems to be a<br />

normal condition of captive living.<br />

CONCLUSIONS<br />

Captive female manatees at different facilities carried out regular non-aggressive<br />

contacts among conspecifics. Aggressive interactions were rare and not associated<br />

with feeding periods. The rates of non-aggressive contacts were density dependent


ut not random. Behaviors were not observed in the first 3 months of captivity nor<br />

were interactions with human handlers documented. Additional studies of captive<br />

and wild manatees are needed to understand association patterns and the role that<br />

relatedness and natal range might play in the level and types of interactions.<br />

ACKNOWLEDGMENTS<br />

We thank Walt Disney World The Living Seas at Epcot Center, Homosassa<br />

Springs Wildlife State Park, Lowry Park Zoo, Miami Seaquarium, Mote Marine<br />

Laboratory, Parker Aquarium in Bradenton and Sea World of Florida for granting<br />

permission to do this research. We also thank the following organizations for helping<br />

support this research: Georgia Southern University’s Academic Excellence Award,<br />

the Graduate Student Professional Development Fund at Georgia Southern<br />

University, and the Jane Smith Turner Foundation and the Eppley Foundation<br />

(awards to B.A.S.). This research was conducted with IACUC approval by Georgia<br />

Southern University and approval from each facility participating in the study.<br />

R. Bonde, J. Koelsch and Drs. R. Chandler, D. Gleason, and L. Leege provided<br />

valuable comments on the manuscript. R. Bonde has approved of the personal<br />

communication reference in the discussion. The research was conducted with the<br />

cooperation of the USGS Sirenia Project (MA79<strong>17</strong>21-2) and we thank their<br />

personnel for cooperation in this study.<br />

REFERENCES<br />

Alexander RD. 1974. The evolution of social<br />

behaviour. Ann Rev Ecol Syst 5:325<strong>–</strong>83.<br />

Beck C, Barros N. 1991. The impact of debris on<br />

the Florida manatee. Mar Poll Bull 22:508<strong>–</strong>10.<br />

Bengtson JL. 1981. Ecology of manatees (Trichechus<br />

manatus) in the St. Johns River, Florida.<br />

[DPhil dissertation]. Twin Cities: University of<br />

Minnesota.<br />

Bossart G, Baden D, Ewing R, Roberts B, Wright<br />

S. 1998. Brevetoxicosis in manatees (Trichechus<br />

manatus latirostris) from the 1996 epizootic:<br />

gross, histologic, and immunohistochemical<br />

features. Toxicol Pathol 26:276<strong>–</strong>82.<br />

Buss IO, Smith, NS. 1966. Observations on<br />

reproduction and breeding behavior of the<br />

African elephant. J Wildl Manage 30:375<strong>–</strong>88.<br />

de Waal FBM. 1995. Bonobo sex and society. Sci<br />

Am 272:82<strong>–</strong>8.<br />

Douglas-Hamilton I. 1972. On the ecology and<br />

behaviour of the African elephant. [DPhil<br />

dissertation]. Oxford: University of Oxford.<br />

Dublin HT. 1983. Cooperation and reproductive<br />

competition among female African elephants.<br />

In: Wasser SK, editor. Social behaviour of<br />

female vertebrates. New York: Academic Press.<br />

p 291<strong>–</strong>313.<br />

Eisenberg JF, McKay GM, Jainudeen, MR. 1971.<br />

Reproductive behavior of the Asiatic elephant<br />

Captive Female Manatee Social Behavior 143<br />

(Elephas maximus maximus L.). Behaviour<br />

38:193<strong>–</strong>225.<br />

Emlen ST. 1997. Predicting family dynamics in<br />

social vertebrates. In: Krebs J, Davies N, editors.<br />

Behavioral ecology, an evolutionary approach.<br />

Cambridge, MA: Blackwell Science. p 228<strong>–</strong>53.<br />

Gadgil M, Nair P. 1983. Observations on the social<br />

behaviour of free ranging groups of tame Asiatic<br />

elephants (Elephas maximus). Anim Sci 93:<br />

225<strong>–</strong>33.<br />

Glickman S, Zabel C, Yoerg S, Weldele M, Drea<br />

C, Frank L. 1997. Social affiliation, and dominance<br />

in the social life of spotted hyenas. In:<br />

Carter C, Lederhendler I, Kirkpatrick B, editors.<br />

The integrative neurobiology of affiliation: animals<br />

of the New York Academy of Science.<br />

Cambridge, MA: MIT Press. p 131<strong>–</strong>9.<br />

Hartman D. 1979. Ecology and behavior of the<br />

manatee (Trichechus manatus) in Florida. Special<br />

publication no. 5. Am Soc Mamm 153 p.<br />

Koelsch J. 1997. The seasonal occurrence and<br />

ecology of Florida manatees (Trichechus manatus<br />

latirostris) in coastal waters near Sarasota,<br />

Florida. [M.S. dissertation]. Sarasota: University<br />

of South Florida.<br />

Lee PC, Moss CJ. 1999. The social context for<br />

learning and behavioural development among<br />

wild African elephants. In: Box HO, Gibson KR,


144 Harper and Schulte<br />

editors. Mammalian social learning: comparative<br />

and ecological perspectives. New York: Cambridge<br />

University Press. p 102<strong>–</strong>25.<br />

Maher CR, Lott DF. 1995. Definitions of territoriality<br />

used in the study of variation in<br />

vertebrate spacing systems. Anim Behav<br />

49:1581<strong>–</strong>97.<br />

Maher CR, Lott DF. 2000. A review of ecological<br />

determinants of territoriality within vertebrate<br />

species. Am Midl Nat 143:1<strong>–</strong>29.<br />

Martin P, Bateson P. 1993. Measuring behaviour:<br />

an introductory guide. Second edition. Cambridge:<br />

Cambridge University Press. 222 p.<br />

McComb K, Moss C, Durant SM, Baker L,<br />

Soyialel S. 2001. Matriarchs as repositories of<br />

social knowledge in African elephants. Science<br />

292:491<strong>–</strong>4.<br />

Moss CJ. 1976. Portraits in the wild: behavior<br />

studies of East African mammals. Chicago (IL):<br />

University of Chicago Press. 363 p.<br />

Pusey A, Packer C. 1999. The ecology of relationships.<br />

In: Krebs J, Davies N, editors. Behavioral<br />

ecology, an evolutionary approach. Cambridge:<br />

Blackwell Science. p 254<strong>–</strong>83.<br />

Reynolds J. 1981. Behavior patterns in the West<br />

Indian manatee, with emphasis on feeding and<br />

diving. Fla Sci 44:233<strong>–</strong>42.<br />

Reynolds J, Odell D. 1991. Manatees and dugongs.<br />

New York: Facts on File, Inc. 192 p.<br />

Schilder MB. 1992. Stability and dynamics of<br />

group composition in a herd of captive plains<br />

zebras. Ethology 90:154<strong>–</strong>68.<br />

Schulte BA. 2000. Social structure and helping<br />

behavior in captive elephants. Zoo Biol 19:447<strong>–</strong>60.<br />

Sokal RR, Rohlf FJ. 1995. Biometry. 3rd edition.<br />

New York (NY): W.H. Freeman and Company.<br />

887 p.<br />

Wells RS, Boness DJ, Rathbun GB. 1999. Behavior.<br />

In: Reynolds JE III, Rommel SA, editors.<br />

Biology of marine mammals. Washington DC:<br />

Smithsonian Institution Press. p 324<strong>–</strong>422.<br />

Wright SD, Ackerman BB, Bonde RK, Beck CA,<br />

Banowetz DJ. 1995. Analysis of watercraftrelated<br />

mortality of manatees in Florida, 1979<strong>–</strong><br />

1991. In: O’Shea T, Ackerman B, Percival H,<br />

editors. Population of biology of the Florida<br />

manatee (Trichechus manatus latirostris).<br />

National Biological Service Information and<br />

Technology Report 1. Washington, DC: US<br />

Department of the Interior. p 259<strong>–</strong>68.<br />

Young JS. 2001. An investigation of captive<br />

Florida manatee (Trichechus manatus latirostris)<br />

behavior and social interactions. [M.S. dissertation].<br />

Statesboro: Georgia Southern University.


Community Structure, Population Control, and Competition<br />

Nelson G. Hairston; Frederick E. Smith; Lawrence B. Slobodkin<br />

The American Naturalist, Vol. 94, No. 879. (Nov. - Dec., 1960), pp. 421-425.<br />

Stable URL:<br />

http://links.jstor.org/sici?sici=0003-0147%28196011%2F12%2994%3A879%3C421%3ACSPCAC%3E2.0.CO%3B2-A<br />

The American Naturalist is currently published by The University of Chicago Press.<br />

Your use of the JSTOR archive indicates your acceptance of JSTOR's Terms and Conditions of Use, available at<br />

http://www.jstor.org/about/terms.html. JSTOR's Terms and Conditions of Use provides, in part, that unless you have obtained<br />

prior permission, you may not download an entire issue of a journal or multiple copies of articles, and you may use content in<br />

the JSTOR archive only for your personal, non-commercial use.<br />

Please contact the publisher regarding any further use of this work. Publisher contact information may be obtained at<br />

http://www.jstor.org/journals/ucpress.html.<br />

Each copy of any part of a JSTOR transmission must contain the same copyright notice that appears on the screen or printed<br />

page of such transmission.<br />

The JSTOR Archive is a trusted digital repository providing for long-term preservation and access to leading academic<br />

journals and scholarly literature from around the world. The Archive is supported by libraries, scholarly societies, publishers,<br />

and foundations. It is an initiative of JSTOR, a not-for-profit organization with a mission to help the scholarly community take<br />

advantage of advances in technology. For more information regarding JSTOR, please contact support@jstor.org.<br />

http://www.jstor.org<br />

Tue Nov 6 13:08:48 2007


Vol. XCIV, No. 879 The American Naturalist November-December, 1960<br />

COMMUNITY STRUCTURE, POPULATION CONTROL,<br />

AND COMPETITION<br />

NELSON G. HAIRSTON, FREDERICK E. SMITH,<br />

AND LARIRENCE B. SLOBODKIN<br />

Department of Zoology, The University of Michigan, Ann Arbor, Michigan<br />

The methods whereby natural populations are limited in size have been<br />

debated with vigor during three decades, particularly during the last few<br />

years (see papers by Nicholson, Birch, Andrewartha, Milne, Reynoldson,<br />

and Hutchinson, and ensuing discussions in the Cold Spring Harbor Symposium,<br />

1957). Few ecologists will deny the importance of the subject,<br />

since the method of regulation of populations must be known before we can<br />

understand nature and predict its behavior-. Although discussion of the subject<br />

has usually been confined to single species populations, it is equally<br />

important in situations where two or more species are involved.<br />

The purpose of this note is to demonstrate a pattern of population control<br />

in many communities which derives easily from a series of general, widely<br />

accepted observations. The logic used is not easily refuted. Furthermore,<br />

the pattern reconciles conflicting interpretations by showing that populations<br />

in different trophic levels are expected to differ in their methods of<br />

control.<br />

Our first observation is that the accumulation of fossil fuels occurs at a<br />

rate that is negligible when compared with the rate of energy fixaeion<br />

through photosynthesis in the biosphere. Apparent exceptions to this observation,<br />

such as bogs and ponds, are successionai stages, in which the<br />

failure of decomposition hastens the termination of the stage. The rate of<br />

accumulation when compared with that of photosynthesis has also been<br />

shown to be negligible over geologic time (Hutchinson, 1948).<br />

If virtually all of the energy fixed in photosynthesis does indeed flow<br />

through he biosphere, it must follow that all organisms taken together are<br />

limited by the amount of energy fixed. In particular, the decomposers as a<br />

group must be food-limited, since by definition they comprise the trophic<br />

level which degrades organic debris. There is no a priori reason why predators,<br />

behavior, physiological changes induced by high densities, etc., could<br />

not limit decomposer populations. In fact, some decomposer populations<br />

may be limited in such ways. If so, however, others must consume the<br />

tt<br />

left-over" food, so that the group as a whole remains food limited; otherwise<br />

fossil fuel would accumulate rapidly.<br />

Any population which is not resource-limited must, of course, be limited<br />

to a level below that set by its resources.<br />

Our next three observations are interrelated. They apply primarily to terrestrial<br />

communities. The first of these is that cases of obvious depletion<br />

of green plants by herbivores are exceptions to the general picture, in which


422 THE AMERICAN NATURALIST<br />

the plants are abundant and largely intact. Moreover, cases of,obvious<br />

mass destruction by meteorological catastrophes are exceptional in most<br />

areas. Taken together, these two observations mean that producers are<br />

neither herbivore-limited nor catastrophe-limited, and must therefore be<br />

limited by their own exhaustion of a resource. In many areas, the limiting<br />

resource is obviously light, but in arid regions water may be the critical<br />

factor, and there are spectacular cases of limitation through the exhaustion<br />

of a critical mineral. The final observation in this group is that there are<br />

temporary exceptions to the general lack of depletion of green plants by<br />

herbivores. This occurs when herbivores are protected either by man or<br />

natural events, and it indicates that the herbivores are able to deplete the<br />

vegetation whenever they become numerous enough, as in the cases of the<br />

Kaibab deer herd, rodent plagues, and many insect outbreaks. It therefore<br />

follows that the usual condition is for populations of herbivores not to be<br />

limited by their food supply.<br />

The vagaries of weather have been suggested as an adequate method of<br />

control for herbivore populations. The best factual clues related to this<br />

argument are to be found in the analysis of the exceptional cases where<br />

terrestrial herbivores have become numerous enough to deplete the vegeta-<br />

tion. This often occurs with introduced rather than native species. It is<br />

most difficult to suppose that a species had been unable to adapt so as to<br />

escape control by the weather to which it was exposed, and at the same<br />

time by sheer chance to be able to escape this control from weather to<br />

which it had not been previously exposed. This assumption is especially<br />

difficult when mutual invasions by different herbivores between two coun-<br />

tries may in both cases result in pests. Even more difficult to accept, how-<br />

ever, is the implication regarding the native herbivores. The assumption<br />

that the hundreds or thousands of species native to a forest have failed to<br />

escape from control by the weather despite long exposure and much selec-<br />

tion, when an invader is able to defoliate without this past history, implies<br />

that "pre-adaptation" is more likely than ordinary adaptation. This we can-<br />

not accept.<br />

The remaining general method of herbivore control is predation (in its<br />

broadest sense, including parasitism, etc.). It is important to note that<br />

this hypothesis is not denied by the presence of introduced pests, since it<br />

is necessary only to suppose that either their natural predators have been<br />

left behind, or that while the herbivore is abIe to exist in the new climate,<br />

its enemies are not. There are, furthermore, numerous examples of the di-<br />

rect effect of predator removal. The history of the Kaibab deer is the best<br />

known example, although deer across the northern portions of the country<br />

are in repeated danger of winter starvation as a result of protection and<br />

predator removal. Several rodent plagues have been attributed to the local<br />

destruction of predators. More recently, the extensive spraying of forests<br />

to kill caterpillars has resulted in outbreaks of scale insects. The latter<br />

are protected from the spray, while their beetle predators and other insect<br />

enemies are not.


POPULATION CONTROL AND COMPETITION<br />

Thus, although rigorous proof that herbivores are generally controlled by<br />

predation is lacking, supporting evidence is available, and the alternate<br />

hypothesis of control by weather leads to false or untenable implications.<br />

The foregoing conclusion has an important implication in the mechanism<br />

of control of the predator populations. The predators and parasites, in con-<br />

trolling the populations of herbivores, must thereby limit their own re-<br />

sources, and as a group they must be food-limited. Although the populations<br />

of some carnivores are obviously limited by territoriality, this kind of in-<br />

ternal check cannot operate for all carnivores taken together. If it did, the<br />

herbivores would normally expand to the point of depletion of the vegeta-<br />

tion, as they do in the absence of their normal predators and parasites.<br />

There thus exists either direct proof or a great preponderance of factual<br />

evidence that in terrestrial communities decomposers, producers, and preda-<br />

tors, as whole trophic levels, are resource-limited in the classical density-<br />

dependent fashion. Each of these three can and does expand toward the<br />

limit of the appropriate resource. We may now examine the reasons why this<br />

is a frequent situation in nature.<br />

Whatever the resource for which a set of terrestrial plant species com-<br />

pete, the competition ultimately expresses itself as competition for space.<br />

A community in which this space is frequently emptied through depletion<br />

by herbivores would run the continual risk of replacement by another assemblage<br />

of species in which the herbivores are held down in numbers by<br />

predation below the level at which they damage the vegetation. That space<br />

once held by a group of terrestrial plant species is not readily given up is<br />

shown by the cases where relict stands exist under climates no longer suit-<br />

able for their return following deliberate or accidental destruction. Hence,<br />

the community in which herbivores are held down in numbers, and in which<br />

the producers are resource-limited will be the most persistent. The develop-<br />

ment of this pattern is less likely where high producer mortalities are in-<br />

evitable. In lakes, for example, algal populations are prone to crash whether<br />

grazed or not. In the same environment, grazing depfetion is much more<br />

common than in communities where the major producers are rooted plants.<br />

A second general conclusion follows from the resource limitation of the<br />

species of three uophic levels. This conclusion is that if more than one<br />

species exists in one of these levels, they may avoid competition only if<br />

each species is limited by factors completely unutilized by any of the other<br />

species. It is a fact, of course, that many species occupy each level in<br />

most communities. It is also a fact that they are not sufficiently segregated<br />

in their needs to escape competition. Although isolated cases of nonoverlap<br />

have been described, this has never been observed for an entire<br />

assemblage. Therefore, interspecific competition for resources exists<br />

among producers, among carnivores, and arllong decomposers.<br />

It is satisfying to note the number of observations that fall into line with<br />

the foregoing deductions. Interspecific competition is a powerful selective<br />

force, and we should expect to find evidence of its operation. Moreover,<br />

the evidence should be most conclusive in trophic levels where it is neces-<br />

423


424 THE AMERICAN NATURALIST<br />

sarily present. Among decomposers we find the most obvious specific<br />

mechanisms for reducing populations of competitors. The abundance of<br />

antibiotic substances attests to the frequency with which these mechanisms<br />

have been developed in the trophic level in which interspecific competition<br />

is inevitable. The producer species are the next most likcly to reveal evi-<br />

dence of competition, and here we find such phenomena as crowding, shad-<br />

ing, and vegetational zonation.<br />

Among the carnivores, however, obvious adaptations for interspecific<br />

competition are less common. Active competition in the form of mutual<br />

habitat-exclusion has been noted in the cases of flatworms (Beauchamp and<br />

Ullyott, 1932) and salamanders (Hairston, 1951). The commonest situation<br />

takes the form of niche diversification as the result of interspecific compe-<br />

tition. This has been noted in birds (Lack, 1945; MacArthur, 1958), sala-<br />

manders (Hairston, 1949), and other groups of carnivores. Quite likely,<br />

host specificity in parasites and parasitoid insects is at least partly due<br />

to the influence of interspecific competition.<br />

Of equal significance is the frequent occurrence among herbivores of ap-<br />

parent exceptions to the influence of density-dependent factors. The grass-<br />

hoppers described by Birch (1957) and the thrips described by Davidson and<br />

Andrewartha (1948) are well known examples. Moreover, it is among herbi-<br />

vores that we find cited examples of coexistence without evidence of com-<br />

petition for resources, such as the leafhoppers reported by Ross (1957), and<br />

the psocids described by Broadhead (19>8). It should be pointed out that<br />

in these latter cases coexistence applies primarily to an identity of food<br />

and place, and other aspects of the niches of these organisms are not known<br />

to be identical.<br />

SUMMARY<br />

In summary, then, our general conclusions are: (1) Populations of producers,<br />

carnivores, and decomposers are limited by their respective resources<br />

in the classical density-dependent fashion. (2) Interspecific com-<br />

petition must necessarily exist among the members of each of these three<br />

trophic levels. (3) Herbivores are seldom food-limited, appear most often<br />

to be predator-limited, and therefore are not likely to compete for common<br />

resources.<br />

LITERATURE CITED<br />

Andrewartha, H. G., 1957, The use of conceptual models in population ecol-<br />

ogy. Cold Spring Harbor Symp. Quant. Biol. 22: 219-232.<br />

Beauchamp, R. S. A., and P. Ullyott, 1937, Competitive relationships be-<br />

tween certain species of fresh-water triclads. J. Ecology 20: 200-<br />

208.<br />

Birch, L. C., 1957, The role of weather in determining the distribution and<br />

abundance of animals. Cold Spring Harbor Symp. Quant. Biol. 22:<br />

2<strong>17</strong>-263.<br />

Broadhead, E., 1958, The psocid fauna of larch trees in northern England.<br />

J. Anim. Ecol. 27: 2<strong>17</strong>-263.


POPULATION CONTROL AND COMPETITION<br />

425<br />

Davidson, J., and H. G. Andrewartha, 1948, The influence of rainfall, evaporation<br />

and atmospheric temperature on fluctuations in the size of a<br />

natural population of Thrzps zmagznis (Thysanoptera). j. him.<br />

Ecol. <strong>17</strong>: 200-222.<br />

Hairston, N. G., 1949, The local distribution and ecology of the Plethodontid<br />

salamanders of the southern Appalachians. Ecol. hlonog.<br />

19: 47-73.<br />

1951, Interspecies competition and its probable influence upon the vertical<br />

distribution of Appalachian salamanders of the genus Plethodon.<br />

Ecology 32: 266-274.<br />

Hutchinson, G. E., 1948, Circular causal systems in ecology. Ann. M.Y.<br />

Acad. Sci. 50: 221-246.<br />

1957, Concluding remarks. Cold Spring Harbor Symp. Quant. Biol. 22:<br />

415-427.<br />

Lack, D., 1945, The ecology of closely related species with special reference<br />

to cormorant (Phalacrocorax carbo) and shag (P, arzstotelzs).<br />

J. Anim, Ecol. 14: 12-16.<br />

MacArthur, R. H., 1958, Population ecology of some warblers of northeastern<br />

coniferous forests. Ecology 39: 599-619.<br />

Milne, A., 1957, Theories of riatural control of insect populations. Cold<br />

Spring Harbor Symp. Quane. Biol. 22: 253-271.<br />

Nicholson, A. J., 1957, The self-adjustment of populations to change, Cold<br />

Spring Harbor Symp. Ouant. Biol. 22: 153-<strong>17</strong>2.<br />

Reynoldson, T. B., 1957, Population fluctuations in Urceolarza rnzt~a (Peritricha)<br />

and Enchytraeus nlbzdus (Oligochaeta) and their bearing on<br />

regulation. Cold Spring Hat bor Symp. Quant. Biol. 22: 313-327.<br />

Ross, H.H., 1957, Principles of natural coexistence indicated by leafhopper<br />

populations. Evolutaon 11 : 113-129.


Environmental Conservation 29 (2): 192<strong>–</strong>206 © 2002 Foundation for Environmental Conservation DOI:10.10<strong>17</strong>/S0376892902000127<br />

The future of seagrass meadows<br />

CARLOS M. DUARTE*<br />

IMEDEA (CSIC <strong>–</strong> UiB), C/ Miquel Marqués 21, 07190 Esporles, (Islas Baleares), Spain<br />

Date submitted: 29 May 2001 Date accepted: 15 February 2002<br />

SUMMARY<br />

Seagrasses cover about 0.1<strong>–</strong>0.2% of the global ocean, and<br />

develop highly productive ecosystems which fulfil a key<br />

role in the coastal ecosystem. Widespread seagrass loss<br />

results from direct human impacts, including mechanical<br />

damage (by dredging, fishing, and anchoring),<br />

eutrophication, aquaculture, siltation, effects of coastal<br />

constructions, and food web alterations; and indirect<br />

human impacts, including negative effects of climate<br />

change (erosion by rising sea level, increased storms,<br />

increased ultraviolet irradiance), as well as from natural<br />

causes, such as cyclones and floods. The present review<br />

summarizes such threats and trends and considers<br />

likely changes to the 2025 time horizon. Present losses<br />

are expected to accelerate, particularly in South-east<br />

Asia and the Caribbean, as human pressure on the<br />

coastal zone grows. Positive human effects include<br />

increased legislation to protect seagrass, increased<br />

protection of coastal ecosystems, and enhanced efforts<br />

to monitor and restore the marine ecosystem. However,<br />

these positive effects are unlikely to balance the negative<br />

impacts, which are expected to be particularly prominent<br />

in developing tropical regions, where the capacity<br />

to implement conservation policies is limited.<br />

Uncertainties as to the present loss rate, derived from<br />

the paucity of coherent monitoring programmes, and<br />

the present inability to formulate reliable predictions as<br />

to the future rate of loss, represent a major barrier to the<br />

formulation of global conservation policies. Three key<br />

actions are needed to ensure the effective conservation<br />

of seagrass ecosystems: (1) the development of a<br />

coherent worldwide monitoring network, (2) the development<br />

of quantitative models predicting the responses<br />

of seagrasses to disturbance, and (3) the education of the<br />

public on the functions of seagrass meadows and the<br />

impacts of human activity.<br />

Keywords: seagrass, conservation, status, perspectives, global<br />

change<br />

INTRODUCTION<br />

Seagrasses ecosystems are widely recognized as key ecosystems<br />

in the coastal zone, with important functions in the<br />

* Correspondence: Professor Carlos M. Duarte Fax: +34 971<br />

61<strong>17</strong>61 e-mail: cduarte@uib.es<br />

marine ecosystem (Hemminga & Duarte 2000). Yet, there is<br />

growing evidence that seagrass meadows are presently experiencing<br />

worldwide decline primarily because of human<br />

disturbance, such as direct physical damage and deterioration<br />

of water quality (Short & Wyllie-Echeverria 1996;<br />

Hemminga & Duarte 2000). There is, therefore, concern that<br />

the functions seagrasses have performed in the marine<br />

ecosystem will be reduced or, in some places, lost altogether.<br />

While efforts are being made to conserve seagrass ecosystems,<br />

these are not guided by a clear forecast of the changes<br />

expected and the threats to come. There is, therefore, a need<br />

for a prospective examination of the expected status of<br />

seagrass ecosystems in the future, which would provide guidance<br />

for the implementation of effective conservation<br />

policies. However, such an exercise requires reliable forecasts<br />

as to the expected progression of the relevant properties<br />

affecting the future status of seagrass ecosystems. Over the<br />

time horizon of 2025, there are estimates of human population<br />

growth and global climate change (Foundation for<br />

Environmental Conservation 2001), and the goal of this<br />

review is to provide a forecast of the likely status of seagrass<br />

ecosystems over that period. However, as Noble laureate Nils<br />

Bohr once stated, ‘prediction is very difficult, particularly if<br />

it concerns the future’, particularly given the many gaps in<br />

the present knowledge on the status of seagrass ecosystems.<br />

Hence, the future scenario depicted here should be<br />

considered as an ecological forecast, rather than a prediction,<br />

driven by the projected trends in the environmental factors<br />

that shape the seagrass habitat. While such forecasts must<br />

necessarily entail considerable uncertainty, their formulation<br />

requires a revision of our understanding of the response of<br />

seagrass ecosystems to a changing ecosystem that may help<br />

identify the areas where more robust knowledge is most<br />

needed. Hence, independently of the reliability of the forecast<br />

issued here, this review will deliver an identification of<br />

the research required to increase the reliability of the predictions<br />

about the response of seagrass ecosystems to a changing<br />

environment.<br />

This review differs from previous attempts to forecast the<br />

future status of seagrass ecosystems in that these focused on<br />

their response to climate change (e.g. Brouns 1994; Beer &<br />

Koch 1996; Short & Neckles 1999), whereas the present exercise<br />

also considers the responses to the pressures derived<br />

from a growing human population. First, seagrasses and the<br />

functions they perform in the marine ecosystem are discussed<br />

together with their response to environmental forcing factors,


including direct and indirect human impacts. Future threats<br />

and the consequences of seagrass loss are used to identify<br />

long-term trends and the likely status of seagrass meadows in<br />

2025. This review concludes by discussing the managerial<br />

frameworks that must be implemented to anticipate the<br />

trends forecasted and optimize the conservation status of<br />

seagrass meadows by year 2025. Indeed, the forecast<br />

provided here reflects a projection from a ‘business as usual’<br />

situation, whereas the adaptive nature of society to change,<br />

through the development of technologies and practices that<br />

minimize human impacts on the coastal environment, may<br />

render the forecasts issued here wrong. The required change<br />

of attitudes and technological developments may be,<br />

however, encouraged by the consideration of the future scenarios<br />

depicted here.<br />

SEAGRASSES AND THEIR FUNCTIONS IN THE<br />

ECOSYSTEM<br />

Seagrasses are angiosperms restricted to growth in marine<br />

environments. These plants evolved early in the history of<br />

angiosperms, but, unlike their land counterparts, showed a<br />

comparative evolutionary stasis (sensu Burt 2001), with their<br />

species richness being unlikely to exceed one hundred species<br />

at any one time through their evolutionary history (den<br />

Hartog 1970; Duarte 2001). All seagrass species are rhizomatous,<br />

clonal plants, occupying space through the reiteration<br />

of shoots, with their leaves and roots produced as a result of<br />

rhizome extension (Marbá & Duarte 1998; Hemminga &<br />

Duarte 2000). This asexual process appears to be the mechanism<br />

for seagrass proliferation, although some species such<br />

as Zostera marina (Olesen 1999) and Enhalus acoroides<br />

(Duarte et al. 1997a) reproduce a lot sexually, a mode of<br />

reproduction which is uncommon in some other species like<br />

Cymodocea serrulata and Posidonia oceanica (Hemminga &<br />

Duarte 2000). Seagrass species vary about 100-fold in growth<br />

rate and lifespan, which is inversely related to size (Duarte<br />

1991a; Marbá & Duarte 1998; Hemminga & Duarte 2000).<br />

Seagrasses require an underwater irradiance generally in<br />

excess of 11% of that incident in the water surface for<br />

growth, a requirement that typically sets their depth limit<br />

(Duarte 1991b). The upslope limit of seagrasses is imposed<br />

by their requirement for sufficient immersion in seawater or<br />

tolerable disturbance by waves and, in northern latitudes, ice<br />

scour (Hemminga & Duarte 2000). Most seagrass species<br />

grow subtidally, although species within some genera such as<br />

Zostera spp., Phyllospadix spp. and Halophila spp. can grow<br />

intertidally (den Hartog 1970; Hemminga & Duarte 2000).<br />

Seagrasses can grow in estuarine and brackish waters, but<br />

require salinity in excess of 5 ‰ to develop (Hemminga &<br />

Duarte 2000). Although some species such as Phyllospadix<br />

spp. grow on rocky shores, most grow on sediments, ranging<br />

from sandy to muddy, with organic contents


194 C.M. Duarte<br />

well as more recent massive losses such as that in Florida Bay<br />

(Robblee et al. 1991), which is one of the largest areas of<br />

seagrass ecosystem worldwide (Fourqurean & Robblee 1999).<br />

The causes and the possible role of human-derived impacts<br />

in such losses are still uncertain (cf. Hemming & Duarte<br />

2000). Strong disturbances, such as damage by hurricanes,<br />

can also lead to major seagrass losses (Preen et al. 1995;<br />

Poiner et al. 1989). Smaller-scale, more recurrent disturbances,<br />

such as that caused by the motion of sand waves in<br />

and out of seagrass patches (Marbá & Duarte 1995) and that<br />

caused by large predators, such as dugongs (Nakaoka & Aioi<br />

1999), represent the main factor structuring some seagrass<br />

landscapes, which are characteristically patchy (Marbá &<br />

Duarte 1995). In contrast, some seagrass species have been<br />

able to form long-lasting meadows, with meadows of the<br />

long-lived Posidonia oceanica dated to >4000 years old<br />

(Mateo et al. 1997), and single clones of Zostera marina dated,<br />

using molecular techniques at 3000 years old (Reusch et al.<br />

1999).<br />

In addition to natural variability of the seagrass habitat,<br />

human intervention is becoming a major source of change to<br />

seagrass ecosystems, whether by direct physical modification<br />

of the habitat by the growing human activity in the coastal<br />

zone (e.g. boating, fishing, construction; Walker et al. 1989),<br />

or through their impact on the quality of waters and sediments<br />

to support seagrass growth (Short & Wyllie-<br />

Echeverria 1996; Hemminga & Duarte 2000), as well as<br />

changes in the marine food webs linked to the seagrasses<br />

(Aragones & Marsh 2000; Jackson 2001; Jackson et al. 2001).<br />

Human impacts<br />

Humans impact seagrass ecosystems, both through direct<br />

proximal impacts, affecting seagrass meadows locally, and<br />

indirect impacts, which may affect seagrass meadows far<br />

away from the sources of the disturbance (Table 1). Proximal<br />

impacts include mechanical damage and damage created by<br />

the construction and maintenance of infrastructures in the<br />

coastal zone, as well as effects of eutrophication, siltation,<br />

coastal engineering and aquaculture (Table 1; Fig. 1).<br />

Indirect impacts include those from global anthropogenic<br />

changes (Fig. 2), such as global warming, sea-level rise, CO 2<br />

and ultraviolet (UV) increase, and anthropogenic impacts on<br />

marine biodiversity, such as the large-scale modification of<br />

the oceanic food web through fisheries ( Jackson 2001;<br />

Jackson et al. 2001). Indirect impacts are already becoming<br />

evident at present (e.g. Beer & Koch 1996).<br />

The most unambiguous source of human impact to<br />

seagrass ecosystems is physical disturbance. This susceptibility<br />

derives from multiple causes, all linked to increasing<br />

human usage of the coastal zone for transportation, recreation<br />

and food production. The coastal zone is becoming an<br />

important focus for services to society, since about 40% of<br />

the human population presently inhabit the coastal zone<br />

(Independent World Commission on the Oceans 1998).<br />

Direct habitat destruction by land reclamation and port<br />

construction is a major source of disturbance to seagrass<br />

meadows, due to dredging and landfill activities, as well as<br />

the reduction in water transparency associated with both<br />

these activities. The construction of new ports is associated<br />

with changes in sediment transport patterns, involving both<br />

increased erosion and sediment accumulation along the adjacent<br />

coast. These changes can exert significant damage on<br />

seagrass ecosystems kilometres away, which can be impacted<br />

by both erosion and burial associated with the changing sedimentary<br />

dynamics (e.g. Pascualini et al. 1999). The operation<br />

of the ports also entails substantial stress to the neighbouring<br />

seagrass meadows, due to reduced transparency and nutrient<br />

Table 1 Impacts of direct and indirect human forcing on seagrass ecosystems.<br />

Type Forcing Possible consequences Mechanisms<br />

Direct impacts Mechanical damage (e.g. trawling, Seagrass loss Mechanical removal and sediment<br />

dredging, push nets, anchoring,<br />

dynamite fishing)<br />

erosion<br />

Eutrophication Seagrass loss Deterioration of light and sediment conditions<br />

Salinity changes Seagrass loss, changes in community Osmotic shock<br />

structure<br />

Shoreline development Seagrass loss due to burial or erosion Seagrass uprooting<br />

Land reclamation Seagrass loss Seagrass burial and shading<br />

Aquaculture Seagrass loss Deterioration of light and sediment conditions<br />

Siltation Seagrass loss and changes in<br />

community structure<br />

Deterioration of light and sediment conditions<br />

Indirect impacts Seawater temperature rise Altered functions and distributions Increased respiration, growth and flowering,<br />

increased microbial metabolism<br />

Increased CO concentration 2 Increased depth limits and Increased photosynthesis, eventual decline of<br />

production calcifying organisms<br />

Sea level rise and shoreline erosion Seagrass loss Seagrass uprooting<br />

Increased wave action and storms Seagrass loss Seagrass uprooting<br />

Food web alterations Changes in community structure Changes in sediment conditions and<br />

disturbance regimes


Nutrient inputs<br />

+<br />

+<br />

+<br />

Phytoplankton<br />

Epiphytic and<br />

opportunistic<br />

algae (+)<br />

- (shading)<br />

(Sediment anoxia)<br />

Grazers<br />

and contaminant inputs associated with ship traffic and<br />

servicing, as well as dredging activities associated with port<br />

and navigation-channel maintenance. Rapid increases in seabased<br />

transport, as well as recreational boating activities have<br />

led to a major increase in the number and size of ports worldwide<br />

(Independent World Commission on the Oceans 1998),<br />

with a parallel increase in the combined disturbance to<br />

seagrass meadows. Ship activity also causes disturbance to<br />

seagrass through anchoring damage, which can be rather<br />

extensive at popular mooring sites (Walker et al. 1989), as<br />

well as fisheries operation, particularly shallow trawling<br />

(Ramos-Esplá et al. 1993; Pascualini et al. 1999) and smallerscale<br />

activities linked to fisheries, such as clam digging and<br />

use of push nets over intertidal and shallow areas and, in<br />

extreme cases, dynamite fishing (Kirkman & Kirkman 2000).<br />

The exponential growth of aquaculture, the fastest-growing<br />

food production industry, has also led to impacts on<br />

seagrasses through shading and physical damage to the<br />

-<br />

-<br />

-<br />

Sediment resuspension<br />

-<br />

Figure 1 Effects of eutrophication derived from increased nutrient<br />

loading on seagrass ecosystems. Positive and negative indicate the<br />

sign of the effects. Chained positive and negative effects affect<br />

seagrasses negatively, whereas chained negative effects affect<br />

seagrasses positively.<br />

New habitat<br />

Sea level rise<br />

+<br />

- Submarine<br />

erosion<br />

deep regression<br />

Increased CO2<br />

Photosynthesis+ - reduced calcification<br />

Seawater warming<br />

Reproduction<br />

habitat<br />

expansion +<br />

-Increased respiration<br />

Figure 2 Forecasted effects of climate change on seagrass<br />

ecosystems. All components include positive and negative effects.<br />

Future of seagrass meadows 195<br />

seagrass beds, as well as deterioration of water and sediment<br />

quality leading to seagrass loss (Delgado et al. 1997, 1999;<br />

Pergent et al. 1999; Cancemi et al. 2000).<br />

The coastal zone also supports increasing infrastructure,<br />

such as pipes and cables for transport of gas, water, energy<br />

and communications, deployment and maintenance of which<br />

also entail disturbance to adjacent seagrass meadows. The<br />

development of coastal tourism, the fastest-growing industry<br />

in the world, has also led to a major transformation of the<br />

coastal zone in areas with pleasant climates. For instance,<br />

about two-thirds of the Mediterranean coastline is urbanized<br />

at the present time, with this fraction exceeding 75% in the<br />

regions with the most developed tourism industry (UNEP<br />

[United Nations Environment Programme] 1989), with<br />

harbours and ports occupying 1250 km of the European<br />

Mediterranean coastline (European Environmental Agency<br />

1999). Urbanization of the coastline often involves destruction<br />

of dunes and sand deposits, promoting beach erosion, a<br />

major problem for beach tourism. Beach erosion, however,<br />

does not only affect the emerged beach, and is usually propagated<br />

to the submarine sand colonized by seagrass,<br />

eventually causing seagrass loss (Medina et al. 2001).<br />

Groynes constructed to prevent beach erosion often create<br />

extensive problems, by altering longshore sediment transport<br />

patterns, further impacting the seagrass ecosystem.<br />

Extraction of marine sand for beach replenishment is only<br />

economically feasible at the shallow depths inhabited by<br />

seagrasses, which are often impacted by these extraction<br />

activities (Medina et al. 2001). The threats coastal tourism<br />

poses to seagrasses are sometimes direct, as in some cases of<br />

purposeful removal of seagrass from beach areas to ‘improve’<br />

beach conditions. Fortunately, there are indications that<br />

coastal tourism is attempting, at least in some areas, to<br />

embrace sustainable principles, including the maintenance of<br />

ecosystem services, such as those provided by seagrasses, and<br />

could well play a role in the future as an agent pressing for<br />

seagrass conservation.<br />

Widespread eutrophication of coastal waters (Vidal et al.<br />

1999) derived from the excessive nutrient input to the sea, is<br />

leading to global deterioration in the quality of coastal waters<br />

(Cloern 2001), which is identified as a major loss factor for<br />

seagrass meadows worldwide (Duarte 1995; Short & Wyllie-<br />

Echeverria 1996; Hemminga & Duarte 2000). Human<br />

activity presently dominates the global nitrogen cycle, with<br />

anthropogenic fixation of atmospheric nitrogen now<br />

exceeding natural sources (Vitousek et al. 1997), and anthropogenic<br />

nitrogen now dominating the reactive nitrogen pools<br />

in the atmosphere, and therefore rainwater, of industrialized<br />

and agricultural areas (Nixon et al. 1996). Hence, anthropogenic<br />

nitrogen dominates the nitrogen inputs to<br />

watersheds, with the human domination of nitrogen fluxes<br />

being reflected in a close relationship between nitrate export<br />

rate and human population in the world’s watersheds (Cole et<br />

al. 1993). Tertiary water-treatment plants only achieve a<br />

partial reduction in nitrogen inputs to the sea, for nitrogen<br />

inputs to the coastal zone are already dominated by direct


196 C.M. Duarte<br />

atmospheric inputs in heavily industrialized or agricultural<br />

areas (Paerl 1995).<br />

Although seagrass meadows are often nutrient limited<br />

(e.g. Duarte 1990), increased nutrient inputs can only be<br />

expected to enhance seagrass primary production at very<br />

moderate levels at best (Borum & Sand-Jensen 1996).<br />

Whereas seagrasses, through their low nutrient requirements<br />

for growth and their high capacity for internal nutrient recycling<br />

(Hemminga & Duarte 2000), are well fitted to cope with<br />

low nutrient availability, other primary producers, both<br />

micro- and macroalgal, are more efficient, because of greater<br />

affinity and higher uptake rates, in using excess nutrient<br />

inputs (Duarte 1995; Hein et al. 1995). Coastal eutrophication<br />

promotes phytoplankton biomass, which deteriorates the<br />

underwater light climate, and the stimulation of the growth<br />

of epiphytes and opportunistic macroalgae, which further<br />

shade and suffocate seagrasses (Duarte 1995; den Hartog<br />

1994; Hemminga & Duarte 2000; Hauxwell et al. 2001). The<br />

alleviation of nutrient limitation, together with the proliferation<br />

of phytoplankton and epiphyte biomass as a result of<br />

increased nutrient inputs imply that coastal eutrophication<br />

leads to a shift from nutrient limitation to light limitation of<br />

ecosystem production (Cloern 2001), enhanced through<br />

competitive interactions between different types of primary<br />

producers for light (Sand-Jensen & Borum 1991; Duarte<br />

1995). These effects result in seagrass loss, particularly in the<br />

deeper portions of the meadows (Sand-Jensen & Borum<br />

1991; Duarte 1995; Borum 1996). Heavy grazing, which can<br />

buffer the negative effects of eutrophication, may attenuate<br />

the effects of overgrowth by phytoplankton, epiphytes and<br />

macroalgae (cf. Heck et al. 2000; Hemminga & Duarte 2000).<br />

Eutrophication may, furthermore, have negative effects<br />

directly derived from the high resulting nutrient concentration,<br />

for high nitrate and ammonium concentrations may<br />

be toxic to seagrasses (e.g. Van Katwijk et al. 1997).<br />

Whereas research on the effects of eutrophication on<br />

seagrass meadows has focused on the effects of reduced light<br />

quality (Duarte 1995), the deterioration of the sediment<br />

conditions may also play a critical role in enhancing the loss<br />

of seagrasses. Seagrass sediments are typically rich in organic<br />

materials, due to the enhanced particle deposition and trapping<br />

under seagrass canopies (Terrados & Duarte 1999;<br />

Gacia & Duarte 2001) compared to adjacent bare sediments.<br />

Microbial processes are, therefore, stimulated in the seagrass<br />

rhizosphere (Hemminga & Duarte 2000), which, if sufficiently<br />

intense, lead to the depletion of oxygen and the<br />

development of bacterial communities with anaerobic metabolism,<br />

which release by-products, such as sulphide and<br />

methane, that may be toxic to seagrasses (e.g. Carlson et al.<br />

1994; Terrados et al. 1999). In order to avoid such toxicity<br />

effects, seagrasses pump a significant fraction of the photosynthetic<br />

oxygen produced to the roots, which release oxygen<br />

to maintain an oxidized microlayer at the root surface<br />

(Pedersen et al. 1998). However, eutrophication reduces<br />

seagrass primary production both through shading and<br />

seagrass loss, thereby reducing the oxygen seagrass roots may<br />

release. This allows anaerobic processes and the resulting<br />

metabolites to accumulate closer to the root surface,<br />

increasing the chances of toxic effects to seagrass. At the same<br />

time, the increased pelagic primary production leads to a<br />

greater input of organic matter to the sediments, enhancing<br />

microbial activity and the sediment oxygen deficit, which<br />

may increase the production of metabolites from anaerobic<br />

microbial metabolism. Both these processes result in the<br />

deterioration of the sediment environment to support<br />

seagrass growth, leading, through its interaction with the<br />

consequences of reduced light availability, to accelerated<br />

seagrass loss (Hemminga & Duarte 2000). Eutrophication<br />

effects on sediments may be more acute where sewage is the<br />

dominant source of nutrients, for this is discharged along<br />

with a high organic load, stimulating microbial activity<br />

( Jorgensen 1996). Aquaculture activities are becoming<br />

increasingly prominent in the shallow, sheltered coastal<br />

waters where seagrass meadows abound. Shading and high<br />

inputs of organic matter from fish cages have been shown to<br />

lead to seagrass decline below and around fish cages (Delgado<br />

et al. 1997, 1999; Pergent et al. 1999, Cancemi et al. 2000),<br />

through processes comparable to those of the eutrophication<br />

outlined above.<br />

Increased siltation of coastal waters is also a major human<br />

impact on seagrass ecosystems, which derives from changes<br />

in land use leading to increased erosion rates and silt export<br />

from watersheds. Siltation is a particularly acute problem in<br />

South-east Asian coastal waters, which receive the highest<br />

sediment delivery in the world as a result of high soil erosion<br />

rates derived from extensive deforestation and other changes<br />

in land use (Miliman & Meade 1993). Siltation severely<br />

impacts South-east Asian seagrass meadows through<br />

increased light attenuation (Bach et al. 1998) and burial<br />

(Duarte et al. 1997b), leading to seagrass loss and, where less<br />

intense siltation occurs, a decline in seagrass diversity,<br />

biomass and production (Terrados et al. 1998).<br />

Large-scale coastal engineering often alters circulation<br />

and salinity distributions, leading to seagrass loss. Hence,<br />

seagrass meadows, previously abundant in Dutch coastal<br />

areas, are now much reduced in surface, partially related to<br />

shifts of coastal waters from marine to brackish or freshwater<br />

regimes (e.g. Nienhuis et al. 1996).<br />

Pollution, other than that from nutrients and organic<br />

inputs, may be an additional source of human impacts on<br />

seagrass ecosystems, although, seagrass appears to be rather<br />

resistant to pollution by organic and heavy metal contaminants<br />

(Hemminga & Duarte 2000). These substances may<br />

possibly harm some components of the seagrass ecosystem,<br />

although such responses have not been examined to a significant<br />

extent.<br />

Human activity in the coastal zone has greatly impacted<br />

biodiversity at regional scales, whether by removing or<br />

adding species. These changes to marine biodiversity impact<br />

food webs and competitive interactions between organisms,<br />

directly and indirectly affecting seagrass beds ( Jackson 2001;<br />

Jackson et al. 2001). Hunting has led to a major decline in the


abundance of dugongs and sea turtles, which are agents of<br />

small-scale disturbance influencing the structure and<br />

dynamics of seagrass ecosystems (e.g. Aragones & Marsh<br />

2000). Indeed, there is concern that the depletion of dugong<br />

and sea turtle populations to levels where they can no longer<br />

exert an impact at the basin scale, may lead to major shifts in<br />

seagrass community structure, with the decline of the species<br />

dependent on recurrent small-scale disturbance, such as a<br />

Halophila spp. (Preen 1995; De Iongh 1996). These concerns<br />

stem from evidence from the fossil record that extinctions<br />

within the one diverse syrenid fauna were associated with<br />

extinctions of seagrass species (Domning 2001). It has been<br />

suggested recently that the decline in mega-grazers may have<br />

caused an increase in organic matter deposition in seagrass<br />

sediments, possibly explaining recent large-scale decline of<br />

seagrasses in systems such as Florida Bay, USA, and<br />

Moreton Bay, Australia ( Jackson 2001; Jackson et al. 2001).<br />

However, the evidence to suggest that the loss of megagrazers<br />

results or has resulted in a deterioration of habitat<br />

conditions to support seagrass is meagre at best ( Jackson<br />

2001; Jackson et al. 2001), so that these suggestions should be<br />

considered as interesting hypotheses pending confirmation,<br />

rather than established facts. In addition, high grazing<br />

pressure in the past, such as implied by this hypothesis,<br />

stresses the plants and ultimately results in a decline of leaf<br />

production, which has been ascribed to a depletion in the<br />

sediment nutrients available for plant growth (Zieman et al.<br />

1984).<br />

Global trade has increased the mobility of marine species,<br />

whether purposely, such as aquarium specimens, or inadvertently,<br />

such as organisms carried in ballast waters. The<br />

increased human-mediated transport of species between<br />

geographically distant locations has increased the incidence<br />

of invasive species. A case affecting Mediterranean, and<br />

probably soon Eastern Pacific seagrasses, is the invasion of<br />

the Mediterranean by the tropical algal species Caulerpa taxifolia,<br />

which first invaded the French Mediterranean in the<br />

early 1980s, apparently released from an aquarium, and has<br />

been reported to have expanded since along the French coast<br />

to reach the Italian and Spanish (Majorca Island) coasts<br />

(Meinesz & Hesse 1991). C. taxifolia grows rapidly and<br />

appears largely to colonize areas devoid of seagrasses, but has<br />

been reported to compete for space and resources with<br />

Posidonia oceanica off Monaco (Meinesz & Hesse 1991), being<br />

able to damage the Posidonia oceanica meadows, particularly<br />

when these are already under stress (de Villéle & Verlaque<br />

1995; Jaubert et al. 1999). The species has recently been<br />

reported on the Californian coast, raising concerns that it<br />

could also cause problems for the seagrass beds there. The<br />

Mediterranean Sea has also been invaded by Halophila stipulacea<br />

(Biliotti & Abdelahad 1990), across the Suez Canal, but<br />

no damage to the local seagrass meadows has been reported.<br />

Direct or indirect human intervention locally causes most<br />

impacts. However, at the regional or global scale, human<br />

activity also exerts an important impact on seagrass ecosystems.<br />

These effects are remarkably difficult to separate from<br />

Future of seagrass meadows 197<br />

responses to background natural changes in the highly<br />

dynamic coastal ecosystem. These impacts involve the effects<br />

of the realized and predicted climate change (Table 1), and<br />

result from changes in sea level, water temperature, UV irradiance,<br />

and CO 2 concentration.<br />

The mean global sea level has risen about 10<strong>–</strong>25 cm over<br />

the 20th century, which should have generated an average<br />

recession of the global coastline by 10<strong>–</strong>25 m (Bruun 1962),<br />

and, therefore, a large-scale erosion of shallow marine sediments.<br />

There is little doubt that such changes must have<br />

affected seagrasses, which are very sensitive to sediment<br />

erosion (e.g. Marbá & Duarte 1995), although there is little<br />

or no direct observational evidence for these changes. The<br />

effects on seagrasses of seawater warming of 0.3<strong>–</strong>0.6 o C over<br />

the 20th century (Mackenzie 1998) are generally less<br />

evident than those produced by sea-level changes.<br />

Temperature affects many processes that determine seagrass<br />

growth and reproduction, including photosynthesis, respiration,<br />

nutrient uptake, flowering and seed germination<br />

(Short & Neckles 1999). Although observations of increased<br />

flowering frequency of Posidonia oceanica in the<br />

Mediterranean have been tentatively linked to the seawater<br />

temperature increase (Francour et al. 1994), there is little<br />

evidence at present to suggest any impact of increased<br />

temperature as a result of global warming. The temperature<br />

increase may further impact seagrass ecosystems through<br />

effects on other components, such as an increased respiratory<br />

rate of the associated microbial communities.<br />

Stimulation of microbial respiration would further enhance<br />

the problems derived from high organic inputs to seagrass<br />

sediments (Terrados et al. 1999). Summer UV irradiance<br />

has greatly increased at high latitudes, and an increase in<br />

the north-temperate zone is also becoming evident.<br />

Increased UV levels are expected to impact negatively<br />

shallow, particularly intertidal, seagrasses (Dawson &<br />

Dennison 1996). The photosynthetic rates of light-saturated<br />

seagrass leaves are often limited by the availability of<br />

dissolved inorganic carbon, and, since the concentration of<br />

CO 2 in well-mixed, shallow coastal waters is in equilibrium<br />

with the atmosphere, the increase in atmospheric CO 2<br />

concentration by 25%, from 290 ppm to 360 ppm, over the<br />

20th century (Mackenzie 1998), may have led to an increase<br />

in light-saturated seagrass photosynthesis by as much as<br />

20% (recalculated from information in Brouns 1994 and<br />

Beer & Koch 1996). There is, however, little evidence that<br />

such physiological responses have led to observable changes<br />

in seagrass ecosystems at present.<br />

Natural versus anthropogenic influences<br />

Some cause-effect relationships between local seagrass loss<br />

and direct human activities, such as increased nutrient and<br />

organic loading, constructions on the coastline or boating<br />

activity, can be readily demonstrated. For instance, the<br />

loss of Zostera marina in a bay on the Atlantic coast of the<br />

USA has been shown to be closely correlated to housing


198 C.M. Duarte<br />

development (Short & Burdick 1996). However, the link<br />

between seagrass losses and indirect human influences is<br />

more elusive, since the coastal zone is a highly dynamic<br />

ecosystem, where many conditions vary simultaneously.<br />

Disturbances such as strong storms, hurricanes and<br />

typhoons, severely impact seagrass beds, to the point that<br />

they may be essential components of the dynamics of seagrass<br />

meadows (Duarte et al. 1997b). The difficulties of discriminating<br />

between sources of seagrass loss are best illustrated by<br />

example. A wasting disease decimated Zostera marina<br />

meadows in the 1930s on both sides of the Atlantic. The<br />

proximal cause for the loss seems to have been an infection by<br />

a fungus, Labyrinthula zosterae, although it may have affected<br />

Zostera marina meadows that were already stressed (den<br />

Hartog 1987). Hypotheses to account for this widespread loss<br />

also point to natural changes, such as unusual seawater<br />

warming, as possible triggers for the decline (den Hartog<br />

1987). Whether indirect human impacts on global processes<br />

may have played a role remains untested. The causes for the<br />

more recent loss of vast areas of seagrass, mostly Thalassia<br />

testudinum, in Florida Bay (Robblee et al. 1991) continue to be<br />

debated, and include, among others, increased anthropogenic<br />

nutrient loading, the effects of climatic changes involving a<br />

long time interval without hurricanes affecting the area also<br />

causing unusually low freshwater discharge (Fourqurean &<br />

Robblee 1999; Zieman et al. 1999), and the effects of the<br />

increased accumulation of detritus derived from loss of<br />

mega-grazers ( Jackson 2001; Jackson et al. 2001). Difficulties<br />

in experimenting at the appropriate scale of whole meadows<br />

to test these hypotheses have precluded elucidation of<br />

whether the decline was due to human or natural causes, or a<br />

combination of both.<br />

The challenge lies, therefore, in separating local loss<br />

processes, which would suggest a local, likely humanderived,<br />

source of disturbance, from indirect effects, both<br />

natural and anthropogenic, acting at a much larger spatial<br />

scale. The ability to reconstruct the growth history of the<br />

long-lived Mediterranean seagrass Posidonia oceanica allowed<br />

the examination of the coherence of past growth patterns<br />

along the Spanish Mediterranean coast (Marbá & Duarte<br />

1997). This examination revealed coherent growth patterns<br />

in populations 200<strong>–</strong>300 km along the coast, as well as a<br />

general tendency towards a decline in vertical growth, which<br />

indicates a tendency towards sediment erosion (Marbá &<br />

Duarte 1997). These trends were present even in meadows<br />

distant from any identifiable human disturbance, and were<br />

related to decadal changes in climate (Marbá & Duarte<br />

1997).<br />

Hence, despite clear signals of anthropogenic effects on<br />

climate components, the responses of the seagrass ecosystem<br />

are still unclear, probably due to the still modest size of the<br />

changes experienced but also, and perhaps to a greater<br />

extent, to the lack of adequate long-term monitoring systems<br />

allowing the detection of responses in the seagrass<br />

ecosystem.<br />

Future threats<br />

Future threats to seagrass meadows evolve from the development<br />

and expansion of those threats already operating.<br />

Increasing human use of the coastal zone predicates an uncertain<br />

future for seagrass meadows. Coastal populations, both<br />

resident and tourist, continue to grow rapidly, with those in<br />

the Mediterranean doubling at 30-year and 15-year intervals,<br />

respectively (UNEP 1989). Increased human activity in the<br />

coastal zone is associated with increased physical disturbance<br />

of seagrass meadows, through the construction of coastal<br />

infrastructures, and increased boating and shipping.<br />

Aquaculture is growing exponentially worldwide (World<br />

Resources Institute 1998), and parallel depletion of natural<br />

fish populations should lead to a maintenance, or even acceleration,<br />

of the expansion of coastal aquaculture, with<br />

associated potential impacts on seagrass meadows. The<br />

present trend towards increasing per caput water use, the<br />

expansion of waterborne sewage systems, and the growing<br />

use of fertilizers (World Resources Institute 1998) should<br />

lead to increasing nutrient and sewage discharge to coastal<br />

waters, and therefore a further expansion of the already widespread<br />

eutrophication problems. Although sewage treatment<br />

is also growing, the volume treated is growing much more<br />

slowly than that of total sewage and nutrient discharge<br />

(Nixon 1995). Similarly, there are no symptoms that land-use<br />

changes and deforestation are declining in the developing<br />

world, so that associated problems, such as siltation of coastal<br />

waters (see above), are likely to continue to increase in the<br />

future.<br />

The extent of indirect human impacts on seagrass ecosystems<br />

should also increase in the future, due to increasing<br />

emissions of CO 2 and other greenhouse gases. These activities<br />

should lead to an increase in the concentration of CO 2 in<br />

seawater, with a resulting decline in seawater pH (e.g.<br />

Kleypas et al. 1999), increased water temperature, and sealevel<br />

rise, together with an increased frequency of storms and<br />

wave action (Carter 1988; Bijlsma et al. 1996). Whereas an<br />

increase in the concentration of CO 2 in seawater may have<br />

positive effects on seagrass photosynthesis, possibly<br />

enhancing seagrass primary production worldwide, the<br />

decline in seawater pH should negatively affect calcifying<br />

organisms (Kleypas et al. 1999), including those present in<br />

seagrass meadows. The increasing sea level and storm activity<br />

(Carter 1988; Bijlsma et al. 1996) should be conducive to<br />

increasing coastal erosion and subsequent loss of seagrass<br />

meadows. Sea-level rise leads to a recession of the coastline,<br />

which can be roughly estimated as a horizontal recession rate<br />

of 100 times the consolidated sea-level rise (applying Bruun’s<br />

1962 rule of 1 m recession per cm increase in sea level).<br />

Sea-level rise creates, in the long term, new habitat for<br />

seagrass growth, which should lead to colonization of the<br />

inundated land, although the rates of sea-level rise are too<br />

slow to allow direct observation of the progression of the<br />

upslope limit of seagrass distribution; however, consideration<br />

of the sea-level rise of 10<strong>–</strong>25 cm over the 20th century


suggests the inundation (applying Bruun’s 1962 rule) of<br />

10<strong>–</strong>25 m of land, with an equivalent potential upslope<br />

progression of the vegetation. The downslope limit of the<br />

seagrass, which is generally set by the water transparency<br />

(Duarte 1991b), is likely to experience a similar upslope<br />

regression (i.e. 1<strong>–</strong>2.5 m horizontally yr -1 ): this assumes the<br />

transparency to remain constant, so that the bathymetric<br />

range occupied by seagrasses should not necessarily increase<br />

with the colonization of new inundated areas upslope. These<br />

processes introduce a dynamic perspective on the upslope<br />

and downslope distribution of seagrass meadows, which<br />

needs be considered in monitoring programmes, which may<br />

interpret the expected horizontal regression of 1<strong>–</strong>2.5 m per<br />

decade at the downslope limit as decline, when it may well<br />

represent a readjustment of the entire meadow to sea-level<br />

rise.<br />

Consequences of seagrass loss<br />

Seagrass loss leads to a loss of the associated functions and<br />

services in the coastal zone. The consequences of seagrass<br />

loss are well documented, through observations of the<br />

changes in the ecosystem upon large-scale seagrass losses<br />

(Hemminga & Duarte 2000). Seagrass loss involves a shift in<br />

the dominance of different primary producers in the coastal<br />

ecosystem, which can only partially compensate for the loss<br />

of primary production. For instance, the increased planktonic<br />

primary production with increasing nutrient inputs<br />

does not compensate for the lost seagrass production, so that<br />

there is no clear relationship between increased nutrient<br />

loading and ecosystem primary production (Borum 1996).<br />

The loss of the sediment protection offered by the seagrass<br />

canopy enhances sediment resuspension, leading to a further<br />

deterioration of light conditions for the remaining seagrass<br />

plants (Olesen 1996). The extent of resuspension can be so<br />

severe following large-scale losses, such as that experienced<br />

during the Zostera marina wasting disease, that the shoreline<br />

may be altered (Christiansen et al. 1981). The loss of<br />

seagrasses will also involve the loss of the oxygenation of sediment<br />

by seagrass roots, promoting anoxic conditions in the<br />

sediments. Seagrass loss has been shown to result in significant<br />

loss of coastal biodiversity, leading to a modification of<br />

food webs and loss of harvestable resources (Young 1978;<br />

Hemminga & Duarte 2000). In summary, seagrass loss represents<br />

a major loss of ecological as well as economic value to<br />

the coastal ecosystems, and is therefore, a major source of<br />

concern for coastal managers.<br />

IDENTIFIED LONG-TERM TRENDS<br />

Whereas records of change in seagrass ecosystems are few,<br />

the existing reports point to a tendency for widespread<br />

decline in both temperate and tropical ecosystems. There<br />

have been reports of large-scale seagrass decline at over 40<br />

locations (Hemminga & Duarte 2000), 70% of which were<br />

Future of seagrass meadows 199<br />

unambiguously attributable to human-induced disturbance<br />

(Short & Wyllie-Echeverria 1996), and reports of seagrass<br />

loss have multiplied by 10 in the last decade, which can only<br />

partially be accounted for by the increased effort in seagrass<br />

research and monitoring (Duarte 1999). Whereas some<br />

recovery has been demonstrated (e.g. Kendrick et al. 1999;<br />

Preen et al. 1995), most records indicate this to be either nonexistent<br />

or incomplete, so that the perceived long-term trend<br />

of seagrass points to a global decline, the extent of which is<br />

largely unknown. Short and Wyllie-Echevarria (1996) estimated,<br />

extrapolating from the reports available, the area of<br />

seagrass meadows globally lost during the 1990s at about<br />

12 000 km 2 , corresponding to the loss during this period alone<br />

of about 2% of the area originally covered by seagrasses.<br />

There is mounting evidence that Posidonia oceanica, which<br />

was estimated to cover about 50 000 km 2 in the<br />

Mediterranean (Bethoux & Copin-Montégut 1986), has been<br />

suffering widespread decline, at least in the north-west<br />

Mediterranean (Marbá et al. 1996; Pascualini et al. 1999;<br />

Boudouresque et al. 2000). An analysis of the age structure of<br />

Posidonia oceanica meadows in the Spanish Mediterranean<br />

showed a tendency for decline in 57% of the meadows examined<br />

(Marbá et al. 1996). Whereas some of those meadows<br />

still persist, some others have already been lost (C.M.<br />

Duarte, unpublished results 1999). The loss is not evenly<br />

distributed along the Spanish Mediterranean, but is concentrated<br />

in approximately 400 km of coastline, where only<br />

traces of some meadows recorded in the late 1980s remain<br />

and many have been lost already (C.M. Duarte, unpublished<br />

results 1999). Whereas local human impacts are often evident<br />

as causes for the decline of Posidonia oceanica meadows, this<br />

is not always the case, and, together with significant correlations<br />

between decadal growth and climate variation (Marbá &<br />

Duarte 1997), points to climate change effects and possibly<br />

indirect human impacts on the observed decline. For<br />

instance, the gradual decline in Zostera marina growing in the<br />

French coast noticed during the 1980s and 1990s was also<br />

related to elevated sea-surface temperatures observed since<br />

the 1980s (Glémarec 1997). These observations suggest that<br />

seagrass decline and, possibly expansion, may be triggered by<br />

large-scale climate change (Fig. 2). If so, rapidly changing<br />

climatic patterns may lead to major changes in seagrass cover<br />

in the future.<br />

Along the Danish coast, eelgrass has shown an important<br />

net reduction in cover over the 20th century, with important<br />

declines (die-off in the 1930s), rebounds and secondary<br />

declines, in between (P.B. Christensen, D. Krause-Jensen &<br />

J. S. Laursen, unpublished data 2000). These observations<br />

portray eelgrass ecosystems as highly dynamic at time scales<br />

longer than decadal. If undisturbed, seagrass communities<br />

can be rather stable, with records of continuous presence of<br />

specific seagrass beds for several millennia (e.g. Zostera<br />

marina, Reusch et al. 1999; Posidonia oceanica, Mateo et<br />

al.1997), and the extent and time scales of fluctuations in<br />

seagrass meadows may differ greatly across species and<br />

locations.


200 C.M. Duarte<br />

POTENTIAL STATUS IN 2025<br />

The prospective trends in both human forcing and climate<br />

change for the next 25 years suggest that the status of seagrass<br />

ecosystems may somewhat improve in some developed countries,<br />

because legislation is being rapidly implemented to<br />

protect seagrass meadows. The zero loss policy in the USA<br />

and Australia (Coles & Fortes 2001) means that any losses<br />

caused by direct human intervention should be compensated<br />

by the creation, generally by transplanting, of a similar extent<br />

of new seagrass meadow. Indeed, the general public is<br />

becoming increasingly aware of the existence of seagrasses<br />

and their important services and functions in coastal ecosystems,<br />

which can only raise the prospect of implementing<br />

conservation management policies further. Tourism, which<br />

has often acted in the past as a vector of degradation of coastal<br />

ecosystems, is shifting towards a more sustainable perspective<br />

and is becoming, in many countries, an agent promoting<br />

an improved quality of coastal ecosystems.<br />

Regrettably, legal protection against seagrass losses is only<br />

possible where disturbance derives from proximal causes,<br />

and difficulties of assigning responsibility for more diffuse<br />

impacts imply that these cannot be corrected through a<br />

posteriori legal actions. This includes eutrophication-derived<br />

seagrass loss, for nutrient inputs are mostly diffuse in nature.<br />

Nutrient reduction plans are being implemented in most<br />

developed countries, although these seem to have had little<br />

impact on the extent of coastal eutrophication, which is likely<br />

to expand in the future, although hopefully at a reduced rate,<br />

despite efforts to reduce nutrient inputs. Moreover, the<br />

trends outlined above only encompass a fraction of the<br />

world’s nations, where losses over the 20th century have<br />

already been large, and include only a modest fraction of the<br />

world’s coastline and seagrass beds.<br />

The bulk of the extant seagrass meadows are found on the<br />

coastline of developing tropical countries, which are experiencing<br />

the greatest rate of environmental degradation at<br />

present, and will continue to do so in the future. The impacts<br />

on seagrass meadows there will derive from land-use changes,<br />

particularly deforestation, and the associated siltation of<br />

coastal waters and growing nutrient and sewage inputs, as<br />

populations, fertilizer use and sewage implementation<br />

expand, and the impacts of the rapidly-growing aquaculture<br />

operations arise. What the loss will be by the year 2025<br />

remains uncertain, since the total extent of these impacts is as<br />

yet unknown. However, the geographic extent of these major<br />

impacts can be readily delimited to the developing world,<br />

particularly affecting seagrass meadows in South-east Asia,<br />

East Africa, and, to a lesser extent, the Caribbean. The<br />

coastal areas of other developing regions, such as South<br />

America and West Africa, support seagrass meadows that are<br />

much more limited in extent, so that the local impacts will<br />

affect smaller areas.<br />

Oceania is the only region with a major seagrass belt where<br />

losses are expected to be relatively small, due to the enforcement<br />

of zero seagrass loss policies and the small population<br />

(29.5 million) relative to the region’s coastline (30 663 km, i.e.<br />

968 humans km -1 ) compared at the other extreme to Asia<br />

(14 197 humans km -1 ; World Resources Institute 1998). Even<br />

so, the coastal zone near Australian cities has already experienced<br />

a significant loss (e.g. Cambridge & McComb 1984;<br />

Clarke & Kirkman 1989), so that seagrasses are considered<br />

amongst the most stressed of Australian ecosystems<br />

(Kirkman 1992; Kirkman & Kirkman 2000).<br />

Seagrass meadows in both developed and developing<br />

regions will be affected by changes in global processes. The<br />

predicted sea-level rise of about 0.5 cm yr -1 (Mackenzie 1998)<br />

implies a further average regression of the global coastline by<br />

12 m by the year 2025 and therefore important submarine<br />

erosion, which should impact seagrass meadows globally.<br />

Quantitative models relating coastline regression to submarine<br />

erosion and seagrass loss are, however, lacking, so that<br />

sensible predictions cannot be formulated. The sea-level rise<br />

may be faster than anticipated due to fast ice melt in polar<br />

regions, land subsidence, and the release of the fresh water<br />

retained in reservoirs reaching the end of their operative<br />

lives. Reservoirs presently hold an amount of water equivalent<br />

to 7 cm of sea-level rise (Carter 1988), and most of them<br />

will have exceeded their operative lives by the year 2025.<br />

This effect may be compensated by construction of new<br />

reservoirs. Submarine erosion may proceed further than<br />

expected because of the predicted increased frequency of<br />

storms and wave energy in the coastal zone with climate<br />

change (Carter 1988; Bijlsma et al. 1996). For similar reasons<br />

the consequences of the predicted increase of 0.5 o C in global<br />

temperature by the year 2025 (Mackenzie 1998) cannot be<br />

quantitatively formulated, although they are likely to be<br />

minor compared to other sources of change. Based on the<br />

calculations in Brouns (1994) and Beer and Koch (1996) the<br />

elevation of CO 2 concentration to almost 400 ppm by the year<br />

2025 may further stimulate seagrass photosynthesis by<br />

10<strong>–</strong>20%. In addition, Zimmerman et al. (1997) calculated<br />

that this increased photosynthesis may increase the depth<br />

limit of seagrasses, which may lead to an expansion in<br />

seagrass cover.<br />

The limitations of knowledge are evident from the fact<br />

that only qualitative predictions on the status of seagrass<br />

ecosystems by year 2025 are possible, because of (1) undefined<br />

rates of change in most relevant forcing factors, (2) lack<br />

of quantitative dose-response models that predict seagrass<br />

response to changes in environmental factors, and (3) absence<br />

of a reliable knowledge of the present extent of seagrass<br />

ecosystems. Even for responses where quantitative predictions<br />

may be formulated, such as the response of seagrass<br />

photosynthesis to increasing CO 2 concentrations, the forecasts<br />

are suspect, for they assume everything else to be held<br />

constant. The predictions therefore lose reliability when the<br />

plants respond in an integrative manner to multiple simultaneous<br />

changes (Chapin et al. 1987), and when these<br />

responses may be affected by responses of other components<br />

of the complex communities that comprise the seagrass<br />

ecosystem. For instance, the positive effect of increased CO 2


concentrations on seagrasses predicted from photosynthetic<br />

rates may be obscured by similar responses by the epiphytes,<br />

buffering the seagrass response due to shading. The<br />

predicted increase in water temperature may enhance<br />

community respiration, resulting in CO 2 concentrations<br />

higher than predicted if in equilibrium with the atmosphere<br />

in some ecosystems. However, enhanced seagrass respiration<br />

may also partially offset the possible increased photosynthetic<br />

gains. One of the major effects of climate change on terrestrial<br />

vegetation is the shift in vegetation ranges with global<br />

warming. A shift in the range of seagrass species, involving a<br />

displacement of subtropical and tropical species towards<br />

higher latitudes, similarly may be expected. This shift should<br />

not affect a poleward extension of the range of temperate<br />

species, for there is no evidence for temperature-dependence<br />

of the latitudinal limit of seagrasses (Hemminga & Duarte<br />

2000; Duarte et al. 2002). In contrast to terrestrial vegetation,<br />

quantitative predictions on the possible shifts in seagrass<br />

ranges with global warming are, however, not possible at the<br />

moment, since these would be dependent on changes in the<br />

oceanic currents and fronts that are themselves difficult to<br />

forecast. The response of seagrass ecosystems to climate<br />

change integrate partial responses at a number of levels, from<br />

physiological to organismal and ecosystem levels, which often<br />

act in opposite directions and interact with one another. The<br />

complex nature of the responses renders our capacity to reliably<br />

predict the response of seagrass ecosystems to climate<br />

change weak.<br />

Provided the recent trends and anticipated threats identified,<br />

it is clear that the most likely scenario is one of major<br />

loss of seagrass ecosystems, although quantitative forecasts<br />

cannot be issued. In a scenario of present and future loss of<br />

seagrass ecosystems, the question of the reversibility of<br />

decline becomes critical, particularly because loss and<br />

recovery of seagrass may be also associated with loss and<br />

recovery of important associated fauna and resources. The<br />

partial recovery of North Atlantic eelgrass populations<br />

affected by the wasting disease within a couple of decades<br />

(Short & Wyllie-Echeverria 1996) shows that recovery from<br />

certain impacts may be possible. However, this is only so<br />

provided that suitable conditions to support seagrass growth<br />

are re-established. Yet, the forcing factors identified above as<br />

likely to cause seagrass decline by year 2025 are forecasted to<br />

increase even further beyond that date so that global-scale<br />

seagrass recovery may not even occur within the present<br />

century. Whereas recovery is possible for species with fast<br />

growth and/or sexual output, such as Halophila spp. and<br />

Enhalus acoroides (Hemminga & Duarte 2000), recovery is<br />

slow in slow-growing species relying largely on clonal expansion,<br />

like the Mediterranean Posidonia oceanica, where<br />

colonization is thought to take place over several centuries<br />

(Duarte 1995). In such species, present and future loss will<br />

therefore affect many generations to come. Recovery may<br />

involve a shift in species composition, towards either fastgrowing,<br />

pioneer species, such as Halophila and Halodule<br />

spp. (Marbá & Duarte 1998), and/or species with high repro-<br />

Future of seagrass meadows 201<br />

ductive outputs, like Enhalus acoroides in the Indo-Pacific<br />

region (Duarte et al. 1997a).<br />

CONCLUSIONS AND MANAGEMENT<br />

Provided the largely qualitative nature of the possible predictions,<br />

the future of seagrass ecosystems remains uncertain.<br />

However, since their present status is already plagued with<br />

problems, and the future global environmental scenario is<br />

one of progressive change, the most likely prospect is a<br />

mounting global seagrass decline throughout the 21st<br />

century. The main threats to seagrass ecosystems on a large<br />

scale will continue to derive from human activity, both<br />

through direct disruption of the coastal zone, changes in land<br />

use and inputs of silt, nutrients and sewage to the coastal<br />

zone, as well as diffuse effects of human activity on climate<br />

(Figs. 1 and 2). Sea-level rise, in particular, is projected to be,<br />

and may already be (e.g. Marbá & Duarte 1997), a major<br />

cause of seagrass decline (Hemminga & Duarte 2000).<br />

Despite this alarm, knowledge of seagrass ecosystems is<br />

still insufficient to even document their possible present and<br />

likely future decline. The 600 000 km2 area global cover of<br />

seagrasses (Charpy-Roubaud & Sournia 1990) is only a best<br />

guess, derived from the application of a few rules of thumb,<br />

in contrast with the relatively adequate knowledge on the area<br />

cover of other key marine ecosystems that can be observed<br />

more easily from space (e.g. mangroves, coral reefs).<br />

Knowledge of the present area seagrasses cover globally is<br />

uncertain, and the situation is probably not much better for<br />

the bulk of the individual nations with an extended coastline.<br />

These uncertainties preclude any reliable quantitative forecast<br />

of future seagrass cover, and only qualitative forecasts<br />

can be formulated based on established trends and expected<br />

changes, supported by anecdotal evidence and opportunistic<br />

observations.<br />

Remote sensing techniques using optical or acoustic<br />

methods allow the monitoring of the surface seagrasses<br />

occupy, allowing the detection of seagrass regression or<br />

expansion (Kendrick et al. 1999; McKenzie et al. 2001) but<br />

cannot resolve changes in the internal density of the vegetated<br />

area. There is therefore a need to improve on the<br />

capacity to monitor seagrass remotely, such as improvements<br />

in acoustic techniques and use of hyperespectral imagerybased<br />

tools to discriminate between seagrass and other<br />

vegetated sediments. Monitoring programmes, typically<br />

based on the assessment of cover and shoot density along<br />

transects and quadrats, generally have an associated error<br />

greater than 30% about the mean (Heidelbaugh & Nelson<br />

1996), making it difficult to reliably detect changes; a reliable<br />

assessment of decline may only be possible for losses already<br />

amounting to as much as 50<strong>–</strong>80%. The low power of monitoring<br />

techniques implies that most monitoring programmes<br />

can only detect a reliable tendency towards seagrass loss<br />

when the seagrass meadows monitored have already experienced<br />

substantial damage. There is, therefore, an urgent need<br />

to design more effective monitoring approaches, capable of


202 C.M. Duarte<br />

detecting losses of 10% or less, as well as to develop early<br />

warning indicators of decline. Repeated census of marked<br />

shoots in plots may allow such precision for some seagrass<br />

species (e.g. Short & Duarte 2001). Use of reconstruction<br />

techniques to assess past seagrass demography (cf. Duarte et<br />

al. 1994) in monitoring programmes has allowed assessment<br />

of local tendencies for decline or expansion and the examination<br />

of the overall demographic balance of very large<br />

seagrass ecosystems (Peterson & Fourqurean 2001). Yet,<br />

even an optimistic analysis clearly indicates that monitoring<br />

efforts cannot possibly encompass but a minimum (< 0.01%)<br />

fraction of the world’s seagrass extent, although the<br />

implementation of seagrass monitoring programmes has<br />

increased over the past decades. The information derived<br />

from individual monitoring efforts could, however, be used<br />

to derive knowledge of trends at larger spatial scales than that<br />

encompassed by the individual monitoring efforts if these<br />

were networked into a coherent programme. The implementation<br />

of coordinated monitoring networks at the national,<br />

regional and global scales would prove instrumental to diagnosis<br />

of the large-scale status and trends of seagrass<br />

ecosystems.<br />

Management practices to address problems affecting<br />

seagrass ecosystems once detected are also insufficiently<br />

developed, for management intervention often fails to revert<br />

an ongoing seagrass decline (e.g. Delgado et al. 1999). The<br />

easiest impacts to regulate are those pertaining to direct<br />

mechanical effects on seagrass ecosystems, as well as some<br />

improvement of seagrass health following regulation of<br />

anchoring in recreational areas (e.g. Cabrera National Park,<br />

Spain; N. Marbá, C.M. Duarte, M. Holmer, R. Martínez, G.<br />

Basterretxea, A. Orfila, A. Jordi & J. Tintoré, personal<br />

communication 2002). Impacts related to changes in water<br />

quality are, however, more difficult to alleviate, although<br />

some successes have been reported, such as the reversal of<br />

seagrass losses following wastewater treatment on the French<br />

Society<br />

- Improved education<br />

- Improved awareness<br />

Managerial<br />

- Monitoring programmes<br />

- Accurate maps<br />

- Impact prevention<br />

- Precautionary principle<br />

- Networking<br />

- Restoration<br />

- Legislation<br />

Mediterranean coast (Pergent-Martini & Pascualini 2000).<br />

Empirical (e.g. Duarte 1991b) aswell as mechanistic (e.g.<br />

Gallegos 1994) models relating water quality to seagrass cover<br />

have been developed, which can be used to build scenarios of<br />

the effectiveness of different management strategies to<br />

improve water quality, and have also led to the development<br />

of metrics based on seagrass performance to assess water<br />

quality (Dennison et al. 1993). Direct management intervention<br />

in the proximal, in addition to the ultimate, causes of<br />

seagrass decline may be possible, such as intervention on<br />

sediment processes. For instance, iron additions to carbonate<br />

sediments may help prevent oxygen consumption by<br />

sulphide oxidation, possibly alleviating seagrass stress from<br />

anoxic sediments with high sulphide concentrations (cf.<br />

Hemminga & Duarte 2000; Chambers et al. 2001). Such<br />

direct intervention practices, however, must be based on<br />

robust knowledge of the proximal causes of seagrass decline,<br />

which requires further research emphasizing experimental<br />

tests of mechanisms and predictive models.<br />

Transplanting techniques to restore seagrasses have had<br />

mixed success (Fonseca 1992; Kirkman 1992), however,<br />

seagrass transplantation is too costly to be implemented at a<br />

large scale, which is a particularly critical drawback provided<br />

the fact that the most important losses are expected in developing<br />

countries. Hence, even though nationwide mangrove<br />

afforestation programmes have been implemented in some<br />

developing countries such as Thailand (Aksornkoae 1993),<br />

similar initiatives for seagrass meadows are as yet not feasible.<br />

In addition, seagrass transplantation programmes often<br />

require damage to a donor population, thereby reducing the<br />

value of transplantation as a management option.<br />

The threats and stresses to seagrass ecosystems are so<br />

diverse that management practices are unlikely to be effective<br />

unless accompanied by changing public attitudes and awareness<br />

(Fig. 3). Hence, the mid- to long-term conservation of<br />

seagrass ecosystems must be based on an adequate education<br />

Improving seagrass ecosystems<br />

Scientific<br />

- Increased knowledge<br />

- Improved predictive<br />

capacity<br />

- Interdisciplinary<br />

approaches<br />

- Improved monitoring<br />

technologies<br />

- Efficient restoration and<br />

intervention techniques<br />

Figure 3 Cooperative elements required to prevent present trend towards seagrass decline and efficiently conserve seagrass ecosystems.


of the public as to the nature of seagrasses, the functions they<br />

perform in nature, and the services these functions provide to<br />

society, and on the formulation and dissemination of best<br />

management practices to conserve seagrass ecosystems, all of<br />

this based on solid, interdisciplinary understanding of<br />

seagrass ecology and how to manage the health of seagrass<br />

ecosystems (Fig. 3). Education campaigns must, therefore,<br />

occupy a prominent place on the agenda for the conservation<br />

of seagrass ecosystems (Kirkman & Kirkman 2000), mobilizing<br />

seagrass scientists, managers and environmental<br />

educators to convey effectively the importance of seagrass<br />

ecosystems and the threats to them. Public awareness alone<br />

cannot be effective unless accompanied by sufficient understanding<br />

of the causes of stress to seagrass, and how these are<br />

integrated at the different levels, from physiological to demographic<br />

and landscape scales, to yield responses to possible<br />

intervention alternatives. In addition, our capacity to predict<br />

seagrass recovery from stress must be developed in order to<br />

allow reliable forecasts to be issued. The future of seagrass<br />

ecosystems will be largely dictated, therefore, by the social<br />

and scientific responses to the challenges ahead.<br />

ACKNOWLEDGEMENTS<br />

The writing of this paper was supported by the European<br />

Commission Managing and Monitoring of Seagrass Beds<br />

project (contract # EVK3-CT-2000-00044). I thank G.<br />

Kendrick and G. Harris for valuable discussion, J. Borum,<br />

D.I. Walker, J. Fourqurean and N.V.C. Polunin for valuable<br />

revisions.<br />

References<br />

Aksornkoae, S. (1993) Ecology and Management of Mangroves.<br />

Bangkok, Thailand: IUCN.<br />

Aragones, L. & Marsh, H. (2000) Impact of dugong grazing and<br />

turtle cropping on tropical seagrass communities. Pacific<br />

Conservation Biology 5: 277<strong>–</strong>288.<br />

Bach, S.S., Borum, J., Fortes, M.D. & Duarte, C.M. (1998) Species<br />

composition and plant performance of mixed seagrass beds along<br />

a siltation gradient at Cape Bolinao, the Philippines. Marine<br />

Ecology Progress Series <strong>17</strong>4: 247<strong>–</strong>256.<br />

Beer, S. & Koch, E. (1996) Photosynthesis of seagrasses vs. marine<br />

macroalgae in globally changing CO2 environments. Marine<br />

Ecology Progress Series 141: 199<strong>–</strong>204.<br />

Bethoux, J.P. & Copin-Montégut, G. (1986) Biological fixation of<br />

atmospheric nitrogen in the Mediterranean Sea. Limnology and<br />

Oceanography 31: 1353<strong>–</strong>8.<br />

Bijlsma, L., Ehler, C.N., Kelin, R.J.L., Kulshrestha, S.M.,<br />

McLean, R.F., Mimumra, N., Nichols, R.J., Nurse, L.A., Pérez<br />

Nieto, H., Stakhiv, E.Z., Turner, K. & Warrick, R.A. (1996)<br />

Coastal zones and small islands. In: Climate Change 1995: Impacts,<br />

Adaptations and Mitigation of Climate Change, ed.<br />

Intergovernmental Panel on Climate Change, pp. 289<strong>–</strong>324.<br />

Cambridge, UK: Cambridge University Press.<br />

Biliotti, A. & Abdelahad, N. (1990) Halophila stipulacea (Frossk.)<br />

Aschers. (Hydrocharticaeae): éspèce nouvelle pour l’Italie.<br />

Posidonia Newsletter 3: 23<strong>–</strong>26.<br />

Future of seagrass meadows 203<br />

Borum, J. (1996) Shallow waters and land/sea boundaries. In:<br />

Eutrophication in Coastal Marine Ecosystems, ed. K. Richardson &<br />

B.B. Jorgensen, pp. <strong>17</strong>9<strong>–</strong>203. Washington DC, USA: American<br />

Geophysical Union.<br />

Borum, J. & Sand-Jensen, K. (1996) Is total primary production in<br />

shallow coastal marine waters stimulated by nitrogen loading?<br />

Oikos 76: 406<strong>–</strong>10.<br />

Boudouresque, C.F., Charbonel, E., Meinesz, A., Pergent, G.,<br />

Pergent-Martini, C., Cadiou, G., Bertrandy, M.C., Foret, P.,<br />

Ragazzi, M. & Rico-Raimondino, V. (2000) A monitoring network<br />

based on the seagrass Posidonia oceanica in the Northwestern<br />

Mediterranean Sea. Biologia Marina Mediterranea 7: 328<strong>–</strong>331.<br />

Brouns, J.J. (1994) Seagrasses and climate change. In: Impacts of<br />

Climate Change on Ecosystems and Species: Marine and Coastal<br />

Systems, ed. J.C. Pernetta, R. Leemans, D. Elder & S. Humphrey,<br />

pp. 59<strong>–</strong>71. Gland, Switzerland: IUCN.<br />

Bruun, P. (1962) Sea level rise as a cause of shore erosion. Journal<br />

Waterways Harbours Division, Proceedings American Society Civil<br />

Engineering 88: 1<strong>17</strong><strong>–</strong>130.<br />

Burt, L. (2001) Evolutionary stasis, contraints, and other terminology<br />

describing evolutionary patterns. Biologial Journal of the<br />

Linnean Society 72: 509<strong>–</strong>5<strong>17</strong>.<br />

Cambridge, M.L. & McComb, A.J. (1984) The loss of seagrasses in<br />

Cockburn Sound, Western Australia. I. The time course and<br />

magnitude of seagrass decline in relation to industrial development.<br />

Aquatic Botany 24: 269<strong>–</strong>285.<br />

Cancemi, G., De Falco, G. & Pergent, G. (2000) Impact of a fish<br />

farming facility on a Posidonia oceanica meadow. Biologia Marina<br />

Mediterranea 7: 341<strong>–</strong>344.<br />

Carlson, P.R., Yabro, L.A. & Barber, T.R. (1994) Relationship of<br />

sediment sulfide to mortality of Thalassia testudinum in Florida<br />

Bay. Marine Science Bulletin 54: 733<strong>–</strong>746.<br />

Carter, R.W.G. (1988) Coastal Environments. London, UK:<br />

Academic Press.<br />

Cebrián, J. & Duarte, C.M. (1997) Patterns in leaf herbivory on<br />

seagrasses. Aquatic Botany 60: 67<strong>–</strong>82.<br />

Chambers, R.M., Fourqurean, J.W., Macko, S.A., Hoppenot, R.<br />

(2001) Biogeochemical effects of iron availability on primary<br />

producers in a shallow marine carbonate sediment. Limnology and<br />

Oceanography 46: 1278<strong>–</strong>1286.<br />

Chapin, F.S., III, Bloom, A.J., Field, C.B. & Waring, R.H. (1987)<br />

Plant responses to multiple environmental factors. BioScience 37:<br />

49<strong>–</strong>87.<br />

Charpy-Roubaud, C. & Sournia, A. (1990) The comparative estimation<br />

of phytoplanktonic and microphytobenthic production in<br />

the oceans. Marine Microbial Food Webs 4: 31<strong>–</strong>57.<br />

Christiansen, C., Christoffersen, H., Dalsgaard, J. & Norberg, R.<br />

(1981) Coastal and nearshore changes correlated with die-back in<br />

eelgrass (Zostera marina L.). Sedimentary Geology 28: 163<strong>–</strong>73.<br />

Clarke, S.M. & Kirkman, H. (1989) Seagrass dynamics. In:<br />

Seagrasses: a Treatise on the Biology of Seagrasses with Special<br />

Reference to the Australian Region, ed. A.W.D. Larkum, A.J.<br />

McComb & S.A. Shepherd, pp. 304<strong>–</strong>34, Amsterdam, the<br />

Netherlands: Elsevier.<br />

Cloern, J.E. (2001) Our evolving conceptual model of the coastal<br />

eutrophication problem. Marine Ecology Progress Series 210:<br />

223<strong>–</strong>253.<br />

Cole, J.J., Peierls, B.L., Caraco, N.F. & Pace, M.L. (1993) Nitrogen<br />

loading of rivers as a human-driven process. In: Humans as<br />

Components of Ecosystems. ed. M.J. McDowell & S.T.A. Rickett<br />

pp. 141<strong>–</strong>157. Springer-Verlag.


204 C.M. Duarte<br />

Coles, R. & Fortes, M. (2001) Protecting seagrass <strong>–</strong> approaches and<br />

methods. In: Global Seagrass Research Methods, ed. F.T. Short &<br />

R.G. Coles, pp. 445<strong>–</strong>463. Amsterdam, the Netherlands: Elsevier.<br />

Costanza, R., d’Arge, R., De Groot, R., Fraber, S., Grasso, M.,<br />

Hannon, B., Limburg, K., Naeem, S., O’Neill, R.V., Paruelo, J.,<br />

Raskin, R.G., Sutton, P. & Van den Belt, M. (1997) The value of<br />

the world’s ecosystem services and natural capital. Nature 387:<br />

253<strong>–</strong>60.<br />

Dawson, S.P. & Dennison, W.C. (1996) Effects of ultraviolet and<br />

photosynthetically active radiation on five seagrass species.<br />

Marine Biology 125: 629<strong>–</strong>638.<br />

De Iongh, H.H. (1996). Plant-herbivore interactions between<br />

seagrasses and dugongs in a tropical small island ecosystem.<br />

Ph.D. Thesis, Catholic University Nijmegen, the Netherlands.<br />

De Villéle, X. & Verlaque, M. (1995) Changes and degradation in a<br />

Posidonia oceanica bed invaded by the introduced tropical alga<br />

Caulerpa taxifolia in the north western Mediterrnean. Botanica<br />

Marina 38: 79<strong>–</strong>87.<br />

Delgado, O., Grau, A., Pou, S., Riera, F., Massuti, C., Zabala, M.<br />

& Ballesteros, E. (1997) Seagrass regression caused by fish<br />

cultures in Fornells Bay, Menorca, western Mediterranean.<br />

Oceanologica Acta 20: 557<strong>–</strong>63.<br />

Delgado, O., Ruiz, J., Pérez, M., Romero, J. & Ballesteros, E. (1999)<br />

Effects of fish farming on seagrass (Posidonia oceanica) in a<br />

Mediterranean bay: seagrass decline after organic loading cessation.<br />

Oceanologica Acta 22: 109<strong>–</strong><strong>17</strong>.<br />

den Hartog, C. (1970) The Seagrasses of the World. Amsterdam, the<br />

Netherlands: North Holland.<br />

den Hartog, C. (1987) ‘Wasting disease’ and other dynamic<br />

phenomena in Zostera beds. Aquatic Botany 27: 3<strong>–</strong>14.<br />

den Hartog, C. (1994) Suffocation of a littoral Zostera bed by<br />

Enteromorpha radiata. Aquatic Botany 47: 21<strong>–</strong>8.<br />

Dennison, W.C., Orth, R.J., Moore, K.A., Stevenson, J.C., Carter,<br />

V., Kollar, S., Bergstrom, P.W. & Batiuk, R. (1993) Assessing<br />

water quality with submersed aquatic vegetation. Bioscience 43:<br />

86<strong>–</strong>91.<br />

Domning, D.P. (2001) <strong>Sirenian</strong>s, seagrasses, and Cenozoic<br />

ecological change in the Caribbean. Palaeogeography,<br />

Palaeoclimatology, Palaeoecology 166: 27<strong>–</strong>50.<br />

Duarte, C.M. (1990) Seagrass nutrient content. Marine Ecology<br />

Progress Series 67: 201<strong>–</strong>207.<br />

Duarte, C.M. (1991a) Allometric scaling of seagrass form and<br />

productivity. Marine Ecology Progress Series 77: 289<strong>–</strong>300.<br />

Duarte, C.M. (1991b) Seagrass depth limits. Aquatic Botany 40:<br />

363<strong>–</strong>377.<br />

Duarte, C.M. (1995) Submerged aquatic vegetation in relation to<br />

different nutrient regimes. Ophelia 41: 87<strong>–</strong>112.<br />

Duarte, C.M. (1999) Seagrass ecology at the turn of the millennium:<br />

challenges for the new century. Aquatic Botany 65: 7<strong>–</strong>20.<br />

Duarte, C.M. (2001) Seagrass ecosystems. In: Encyclopedia of<br />

Biodiversity, Volume 5, ed. S.L. Levin, pp. 255-268. San Diego,<br />

USA: Academic Press.<br />

Duarte, C.M. & Cebrián, J. (1996) The fate of marine autotrophic<br />

production. Limnology and Oceanography 41: <strong>17</strong>58<strong>–</strong>66.<br />

Duarte, C.M. & Chiscano, C.L. (1999) Seagrass biomass and<br />

production: a reassessment. Aquatic Botany 65: 159<strong>–</strong><strong>17</strong>4.<br />

Duarte, C.M., Marbà, N., Agawin, N.S.R., Cebrián, J., Enríquez,<br />

S., Fortes, M.D., Gallegos, M.E., Merino, M., Olesen, B., Sand-<br />

Jensen, K., Uri, J. & Vermaat, J. (1994) Reconstruction of<br />

seagrass dynamics: age determinations and associated tools for<br />

the seagrass ecologist. Marine Ecology Progress Series 107:<br />

195<strong>–</strong>209.<br />

Duarte, C.M., Martínez, R. & Barrón, C. (2002) Biomass, production<br />

and rhizome growth near the northern limit of seagrass<br />

(Zostera marina L.) distribution. Aquatic Botany (in press).<br />

Duarte, C.M., Uri, J., Agawin, N.S.R., Fortes, M.D., Vermaat, J.E.<br />

& Marbá, N. (1997a) Flowering frequency of Philippine<br />

seagrasses. Botanica Marina 40: 497<strong>–</strong>500.<br />

Duarte, C.M., Terrados, J., Agawin, N.S.W., Fortes, M.D., Bach,<br />

S. & Kenworthy, W.J. (1997b) Response of a mixed Philippine<br />

seagrass meadow to experimental burial. Marine Ecology Progress<br />

Series 147: 285<strong>–</strong>294.<br />

Enríquez S., Duarte, C.M. & Sand-Jensen, K. (1993) Patterns in<br />

decomposition rates among photosynthetic organisms: the<br />

importance of detritus C:N:P content. Oecologia 94: 457<strong>–</strong>471.<br />

European Environmental Agency (1999) State and Pressures of the<br />

Marine and Coastal Mediterranean Environment. Luxembourg:<br />

European Environmental Agency.<br />

Fonseca, M.S. (1992) Restoring seagrass systems in the United<br />

States. In: Restoring the Nation’s Marine Environment, ed. G.W.<br />

Thayer, pp. 79<strong>–</strong>110. College Park, Maryland, USA: Maryland<br />

Sea Grant College.<br />

Foundation for Environmental Conservation (2001) An introduction<br />

to long-term environmental trends [WWW document]. URL<br />

http://www.ncl.ac.uk/icef<br />

Fourqurean, J.W. & Robblee, M.B. (1999) Florida Bay: a history of<br />

recent ecological changes. Estuaries 22: 345<strong>–</strong>357.<br />

Francour, P., Bouderesque, C.F., Harmelin, J.G., Harmelin-<br />

Vivien, M.L. & Quignard, J.P. (1994) Are the Mediterranean<br />

waters becoming warmer: information from biological indicators.<br />

Marine Pollution Bulletin 28: 523<strong>–</strong>526.<br />

Gacia, E. & Duarte, C.M. (2001) Elucidating sediment retention by<br />

seagrasses: sediment deposition and resuspension in a<br />

Mediterranean (Posidonia oceanica) meadow. Estuarine Coastal<br />

and Shelf Science 52: 505<strong>–</strong>514.<br />

Gallegos, C.L. (1994) Refining habitat requirements of submersed<br />

aquatic vegetation: role of optical models. Estuaries <strong>17</strong>: 187<strong>–</strong>199.<br />

Glémarec, M. (1979) Les fluctuations temporelles des peuplements<br />

benthiques liées aux fluctuations climatiques. Oceanologica Acta 2:<br />

365<strong>–</strong>71.<br />

Harrison, P.G. (1989) Detrital processing in seagrass systems: a<br />

review of factors affecting decay rates, remineralization and detritivory.<br />

Aquatic Botany 23: 263<strong>–</strong>288.<br />

Hauxwell, J., Cebrián, J., Furlong, C., Valiela, I. (2001) Macroalgal<br />

canopies contribute to eelgrass (Zostera marina) decline in<br />

temperate estuarine ecosystems. Ecology 82: 1007<strong>–</strong>1022.<br />

Heck, K.L., Pennock, J.R., Valentine, J.F., Cohen, L.D. & Sklenar,<br />

S.A. (2000) Effects of nutrient enrichment and small predator<br />

density on seagrass ecosystems: an experimental assessment.<br />

Limnology Oceanography 45: 1041<strong>–</strong>1057.<br />

Heidelbaugh, W.S. & Nelson, W.G. (1996) A power analsis of<br />

methods for assessment of change in seagrass cover. Aquatic<br />

Botany 53: 227<strong>–</strong>233.<br />

Hein, M., Pedersen, M.F. & Sand-Jensen, K. (1995) Size-dependent<br />

nitrogen uptake in micro- and macroalgae. Marine Ecology<br />

Progress Series 118: 247<strong>–</strong>253.<br />

Hemminga, M. & Duarte, C.M. (2000) Seagrass Ecology.<br />

Cambridge, UK: Cambridge University Press.<br />

Independent World Commission on the Oceans (1998) The Ocean.<br />

Our Future. Cambridge, UK: Cambridge University Press.<br />

Jackson, J.B.C. (2001) What was natural in the coastal ocean?<br />

Proceedings of the National Academy of Sciences USA 98:<br />

5411<strong>–</strong>5418.<br />

Jackson, J.B.C., Kirby, M.X., Berger, W.H., Bjorndal, K.A.,


Botsford, L.W., Bourque, B.J., Bradbury, R.H., Cooke, R.,<br />

Erlandson, J., Estes, J.A., Hughes, T.P., Kidewell, S., Lange,<br />

C.B., Lenihan, H.S., Pandolfi, J.M., Peterson, C.H.,<br />

Steneck, R.S., Tegner, M.J. & Warner, R. (2001) Historical overfishing<br />

and the recent collapse of coastal ecosystems. Science 293:<br />

629<strong>–</strong>638.<br />

Jaubert, J.M., Chisholm, J.R.M., Ducrort, D., Ripley, H.T., Roy,<br />

L. & Passeron-Seitre, G. (1999) No deleterious alterations in<br />

Posidonia beds in the Bay of Menton (France) eight years after<br />

Caulerpa taxifolia colonization. Journal of Phycology 35:<br />

1113<strong>–</strong>1119.<br />

Jorgensen, B.B. (1996) Material flux in the sediment. In:<br />

Eutrophication in Coastal Marine Ecosystems, ed. K. Richardson &<br />

B.B. Jorgensen, pp. 115<strong>–</strong>135. Washington DC, USA: American<br />

Geophysical Union.<br />

Kendrick, G.A., Eckersley, J. & Walker, D.I. (1999) Landscapescale<br />

changes in seagrass distribution over time: a case study from<br />

Success Bank, Western Australia. Aquatic Botany 65: 293<strong>–</strong>309.<br />

Kirkman, H. (1992) Large-scale restoration of seagrass meadows.<br />

In: Restoring the Nation’s Marine Environment, ed. G.W. Thayer,<br />

pp. 111<strong>–</strong>140. College Park, Maryland, USA: Maryland Sea Grant<br />

College.<br />

Kirkman, H. & Kirkman, J.A. (2000) The management of seagrasses<br />

in South East Asia and Australia. Biologia Marina Mediterranea 7:<br />

305<strong>–</strong>319.<br />

Kleypas, J. A., Buddemeier, R. W., Archer, D., Gattuso, J.-P.,<br />

Langdon, C. & Opdyke, B. N. (1999) Geochemical consequences<br />

of increased atmospheric CO 2 on coral reefs. Science 284:<br />

118<strong>–</strong>120.<br />

Koch, E.W. (2001) Beyond light: physical, geological and geochemical<br />

parameters as possible submersed aquatic vegetation habitat<br />

requirements. Estuaries 24: 1<strong>–</strong><strong>17</strong>.<br />

Mackenzie, F.T. (1998) Our Changing Planet. Second edition.<br />

Upper Saddle River, NJ, USA: Prentice Hall.<br />

Marbà, N. & Duarte, C.M. (1995) Coupling of seagrass (Cymodocea<br />

nodosa) patch dynamics to subaqueous dune migration. Journal of<br />

Ecology 83: 381<strong>–</strong>389.<br />

Marbá, N. & Duarte, C.M. (1997) Interannual changes in seagrass<br />

(Posidonia oceanica) growth and environmental change in the<br />

Spanish Mediterranean littoral. Limnology and Oceanography 42:<br />

800<strong>–</strong>810.<br />

Marbá, N. & Duarte, C.M. (1998) Rhizome elongation and seagrass<br />

clonal growth. Marine Ecology Progress Series <strong>17</strong>4: 269<strong>–</strong>280.<br />

Marbá, N., Duarte, C.M., Cebrián, J., Enríquez, S., Gallegos,<br />

M.E., Olesen, B. & Sand-Jensen, K. (1996) Growth and population<br />

dynamics of Posidonia oceanica on the Spanish<br />

Mediterranean coast: elucidating seagrass decline. Marine Ecology<br />

Progress Series 137: 203<strong>–</strong>213.<br />

Mateo, M.A., Romero, J., Pérez, M., Littler, M.M. & Littler, D.S.<br />

(1997) Dynamics of millenary organic deposits resulting from the<br />

growth of the Mediterranean seagrass Posidonia oceanica.<br />

Estuarine Coastal Shelf Science 44: 103<strong>–</strong>110.<br />

McKenzie, L.J., Finkbeiner, M.A. & Kirkman, H. (2001) Methods<br />

for mapping seagrass distribution. In: Global Seagrass Research<br />

Methods, ed. F.T. Short & R.G. Coles, pp. 101<strong>–</strong>121. Amsterdam,<br />

the Netherlands: Elsevier.<br />

Medina, J.R., Tintoré, J. & Duarte, C.M. (2001) Las praderas de<br />

Posidonia y la regeneración de playas. Revista de Obras Públicas<br />

3409: 31<strong>–</strong>43.<br />

Meinesz, A. & Hesse, B. (1991) Introduction et invasion de l’algue<br />

Caulerpa taxifolia en Méditerranée nord-occidentale.<br />

Oceanologica Acta 14: 415<strong>–</strong>426.<br />

Future of seagrass meadows 205<br />

Menzies, R.J., Zaneveld, J.S. & Pratt, R.M. (1967) Transported<br />

turtle grass as a source of organic enrichment of abyssal sediments<br />

off North Carolina. Deep-Sea Research 14: 111<strong>–</strong>2.<br />

Miliman, J.D. & Meade, R.H. (1993) World-wide delivery of river<br />

sediment to the oceans. Journal of Geology 91: 1<strong>–</strong>21.<br />

Nakaoka, M. & Aioi, K. (1999) Growth of the seagrass Halophila<br />

ovalis at dugong trails compared to existing within-patch variation<br />

in a Thailand intertidal flat. Marine Ecology Progress Series<br />

184: 97<strong>–</strong>103.<br />

Nixon, S.W. (1995) Coastal marine eutrophication: a definition,<br />

social causes, and future concerns. Ophelia 41: 199<strong>–</strong>219.<br />

Nixon, S.W., Ammerman, J., Atkinson, L., Berounsky, V., Billen,<br />

G., Boicourt, W., Boynton, W., Church, T., DiToro, D.,<br />

Elmgren, R., Garber, J., Giblin, A., Jahnke, R., Owens, N.,<br />

Pilson, M.E.Q. & Seitzinger, S. (1996) The fate of nitrogen and<br />

phosphorus at the land-sea margin of the North Atlantic Ocean.<br />

Biogeochemistry 35: 141<strong>–</strong>180.<br />

Nienhuis, P.H., De Bree, B.H.H., Herman, P.M.J., Holland,<br />

A.M.B., Verschuure, J.M. & Wessel, E.G.J. (1996) Twenty five<br />

years of changes in the distribution and biomass of eelgrass,<br />

Zostera marina, in Grevelingen Lagoon, The Netherlands.<br />

Netherlands Journal of Aquatic Ecology 30: 107<strong>–</strong><strong>17</strong>.<br />

Ochieng, C.A. & Erftemeijer, P.L.A. (1999) Accumulation of<br />

seagrass beach cast along the Kenyan coast: a quantitative assessment.<br />

Aquatic Botany 65: 221<strong>–</strong>238.<br />

Olesen, B. (1996) Regulation of light attenuation and eelgrass<br />

Zostera marina depth distribution in a Danish embayment. Marine<br />

Ecology Progress Series 134: 187<strong>–</strong>94.<br />

Olesen, B. (1999) Reproduction in Danish eelgrass (Zostera marina<br />

L.) stands: size-dependence and biomass partitioning. Aquatic<br />

Botany 65: 209<strong>–</strong>219.<br />

Paerl, H.W. (1995) Coastal eutrophication in relation to atmospheric<br />

nitrogen deposition: current perspectives. Ophelia 41:<br />

237<strong>–</strong>259.<br />

Pascualini, V., Pergent-Martini, C. & Pergent, G. (1999)<br />

Environmental impact identification along the Corsican coast<br />

(Mediterranean sea) using image processing. Aquatic Botany 65:<br />

311<strong>–</strong>320.<br />

Pedersen, O., Borum, J., Duarte, C.M. & Fortes, M.D. (1998)<br />

Oxygen dynamics in the rhizosphere of Cymodocea rotundata.<br />

Marine Ecology Progress Series 169: 283<strong>–</strong>288.<br />

Pergent-Martini, C. & Pascualini, V. (2000) Seagrass population<br />

dynamics before and after the setting up of a wastewater treatment<br />

plant. Biologia Marina Mediterranea 7: 405<strong>–</strong>408.<br />

Pergent, G., Mendez, S., Perget-Martini, C. & Pascualini, V. (1999)<br />

Preliminary data on the impact of fish farming facilities on<br />

Posidonia oceanica meadows in the Mediterranean. Oceanologica<br />

Acta 22: 95<strong>–</strong>107.<br />

Peterson, B.J. & J.W. Fourqurean (2001) Large-scale patterns in<br />

seagrass (Thalassia testudinum) demographics in south Florida.<br />

Limnology and Oceanography 46: 1077<strong>–</strong>1090.<br />

Poiner, I.R., Walker, D.I. & Coles, R.G. (1989) Regional studies <strong>–</strong><br />

seagrasses of tropical Australia. In: Biology of Seagrasses, ed.<br />

A.W.D Larkum, A.J. McComb & S.A. Shepherd, pp. 279<strong>–</strong>303.<br />

Amsterdam, the Netherlands: Elsevier.<br />

Preen, A. (1995) Impacts of dugong foraging on seagrass habitats:<br />

observational and experimental evidence for cultivation grazing.<br />

Marine Ecology Progress Series 124: 201<strong>–</strong>13.<br />

Preen, A.R., Lee Long, W.J. & Coles, R.G. (1995) Flood and<br />

cyclone related loss, and partial recovery, of more than 1000 km 2<br />

of seagrass in Hervey Bay, Queensland, Australia. Aquatic Botany<br />

52: 3<strong>–</strong><strong>17</strong>.


206 C.M. Duarte<br />

Ramos-Esplá, A., Martínez-Pérez, L., Aranda, A., Guillen, J.E.,<br />

Sánchez-Jerez, P. & Sánchez-Lizaso, J.L. (1993) Protección de la<br />

praderas de Posidonia oceanica (L.) Delile mediante arrecifes artificiales<br />

disuasorios frente a la pesca de arrastre ilegal: el caso de El<br />

Campello (SE Ibérico). Publicaciones Especiales Instituto Español<br />

Oceanografía 11: 431<strong>–</strong>439.<br />

Reusch, T.B.H., Borström, C., Stam, W.T. & Olsen, J.L. (1999) An<br />

ancient eelgrass clone in the Baltic. Marine Ecology Progress Series<br />

183: 301<strong>–</strong>304.<br />

Robblee, M.B., Barber, T.R., Carlson, P.R. Jr, Durako, M.J.,<br />

Fourqurean, J.W., Muehlstein, L.K., Porter, D., Yarbro, L.A.,<br />

Zieman, R.T. & Zieman, J.C. (1991) Mass mortality of the tropical<br />

seagrass Thalassia testudinum in Florida Bay (USA). Marine<br />

Ecology Progress Series 71: 297<strong>–</strong>9.<br />

Sand-Jensen, K. & Borum, J. (1991) Interactions among phytoplankton,<br />

periphyton, and macrophytes in temperate freshwaters<br />

and estuaries. Aquatic Botany 41: 137<strong>–</strong>76.<br />

Short, F.T. & Burdick, D.M. (1996) Quantifying seagrass habitat<br />

loss in relation to housing development and nitrogen loading in<br />

Waquoit Bay, Massachusetts. Estuaries 19: 730<strong>–</strong>739.<br />

Short, F.T. & Duarte, C.M. (2001) Methods for the measurement<br />

of seagrass growth and production. In: Global Seagrass Research<br />

Methods, ed. F.T. Short & R.G. Coles, pp: 155-182. Amsterdam,<br />

the Netherlands: Elsevier.<br />

Short, F.T. & Neckles, H.A. (1999) The effects of global climate<br />

change on seagrasses. Aquatic Botany 63: 169<strong>–</strong>196.<br />

Short, F.T. & Wyllie-Echeverria, S. (1996) Natural and humaninduced<br />

disturbance of seagrasses. Environmental Conservation 23:<br />

<strong>17</strong><strong>–</strong>27.<br />

Terrados, J. & Duarte, C.M. (1999) Experimental evidence of<br />

reduced particle resuspension within a seagrass (Posidonia<br />

oceanica L.) meadow. Journal of Experimental Marine Biology and<br />

Ecology 243: 45<strong>–</strong>53.<br />

Terrados, J., Duarte, C.M., Fortes, M.D., Borum, J., Agawin,<br />

N.S.R., Bach, S., Thampanya, U., Kamp-Nielsen, L.,<br />

Kenworthy, W.J., Geertz-Hansen, O. & Vermaat, J. (1998)<br />

Changes in community structure and biomass of seagrass<br />

communities along gradients of siltation in SE Asia. Estuarine,<br />

Coastal and Shelf Science 46: 757<strong>–</strong>768.<br />

Terrados, J., Duarte, C.M., Kamp-Nielsen, L., Borum, J., Agawin,<br />

N.S.R., Fortes, M.D., Gacia, E., Lacap, D., Lubanski, M. &<br />

Greve, T. (1999) Are seagrass growth and survival affected by<br />

reducing conditions in the sediment? Aquatic Botany 65: <strong>17</strong>5<strong>–</strong>197.<br />

UNEP (1989) State of the Mediterranean Marine Environment.<br />

Athens, Greece: MAP Technical Report Series, Volume 28.<br />

Van Katwijk, M.M., Vergeer, L.H.T., Schmitz, G.H.W. & Roelofs,<br />

J.G.M. (1997) Ammonium toxicity in eelgrass Zostera marina.<br />

Marine Ecology Progress Series 157: 159<strong>–</strong>73.<br />

Vidal, M., Duarte, C.M. & Sánchez, M.C. (1999) Coastal eutrophication<br />

research in Europe: progress and imbalances. Marine<br />

Pollution Bulletin 38: 851<strong>–</strong>854.<br />

Vitousek, P.M., Mooney, H.A., Lubchenco, J. & Melillo, J.M.<br />

(1997) Human domination of Earth’s ecosystems. Science 277:<br />

494<strong>–</strong>499.<br />

Walker, D.I., Lukatelich, R.J., Bastyan, G. & McComb, A.J. (1989)<br />

Effect of boat moorings on seagrass beds near Perth, Western<br />

Australia. Aquatic Botany 36: 69<strong>–</strong>77.<br />

World Resources Institute (1998) World Resources 1998<strong>–</strong>99. Oxford,<br />

UK: Oxford University Press: 369 pp.<br />

Young, P.C. (1978) Moreton Bay, Queensland: a nursey area for<br />

juvenile penaeid prawns. Australian Journal Marine Freshwater<br />

Research 29: 55<strong>–</strong>57.<br />

Zieman, J.C., Fourqurean, J.W., Frankovich, T.A. (1999) Seagrass<br />

dieoff in Florida Bay (USA): long-term trends in abundance and<br />

growth of turtle grass, Thalassia testudinum. Estuaries 22:<br />

460<strong>–</strong>470.<br />

Zieman, J.C., Iverson, R.L. & Ogden, J.C. (1984). Herbivory effects<br />

on Thalassia testudinum leaf growth and nitrogen content. Marine<br />

Ecology Progress Series 15: 151<strong>–</strong>8.<br />

Zimmerman, R.C., Kohrs, D.G., Steller, D.L. & Alberte, R.S.<br />

1997. Impacts of CO 2 enrichment on productivity and light<br />

requirements of eelgrass. Plant Physiology 115: 599<strong>–</strong>607.


OUTLOOK ON EVOLUTION AND SOCIETY<br />

doi:10.1111/j.1558-5646.2009.00911.x<br />

IS THE AGE OF THE EARTH ONE OF OUR<br />

“SOREST TROUBLES?” STUDENTS’<br />

PERCEPTIONS ABOUT DEEP TIME AFFECT<br />

THEIR ACCEPTANCE OF EVOLUTIONARY<br />

THEORY<br />

Sehoya Cotner, 1,2 D. Christopher Brooks, 3 and Randy Moore 1<br />

1 Biology Program, University of Minnesota, Minneapolis, Minnesota 55455<br />

2 E-mail: harri054@umn.edu<br />

3 Office of Information Technology, University of Minnesota, Minneapolis, Minnesota 55455<br />

Received July 20, 2009<br />

Accepted November 10, 2009<br />

KEY WORDS: Age of the Earth, evolution education, science education.<br />

When Charles Darwin was developing his ideas for On the Origin<br />

of Species, the most widely accepted estimates of Earth’s<br />

age were those of William Thomson (later Lord Kelvin). Kelvin<br />

used calculations involving thermodynamics to argue that Earth is<br />

only 20<strong>–</strong>100 million years old—an age far too brief to accommodate<br />

evolution by natural selection. Darwin referred to Thomson’s<br />

claim as one of his “sorest troubles,” for Darwin understood that<br />

the history of life on Earth ultimately relies on geology. Darwin<br />

suspected that Earth was much older than Thomson claimed, but<br />

Thomson’s enormous stature as a scientist obliged Darwin to reconcile<br />

his claims with Kelvin’s data. To accommodate Kelvin’s<br />

timeline, Darwin proposed pangenesis as an explanation of inheritance<br />

(i.e., every sperm and egg contained “gemmules thrown off<br />

from each different unit throughout the body”). Darwin’s explanation<br />

sped evolution while avoiding Lamarck’s quasi-spiritual<br />

sources of acquired traits. However, Darwin’s explanation of inheritance<br />

was wrong (see discussion in Moore et al. 2009a).<br />

The age of Earth remains a divisive topic in the modern<br />

evolution<strong>–</strong>creationism controversy. Whereas mainstream science<br />

has long acknowledged that Earth is approximately 4.5 billion<br />

years old, a vocal group of citizens and religious activists con-<br />

858<br />

tinue to insist that Earth is less than 10,000 years old. Although<br />

most geocentrists and flat-Earth advocates have capitulated to<br />

scientific evidence, young-Earth creationists continue to reject<br />

scientific evidence in favor of religious dictum to claim that Earth<br />

is less than 10,000 years old. These antiscience claims have been<br />

surprisingly popular with the public. For example, a Gallup Poll<br />

in early 2009 reported that “On Darwin’s [200th] Birthday, Only<br />

4 in 10 Believe in Evolution” (Newport 2009), and Berkman et al.<br />

(2008) noted that “16% [of biology teachers] believed that human<br />

beings were created by God in their present form at one time<br />

within the last 10,000 years.” In another study, 12.5% of students<br />

were young-Earth creationists (Rutledge and Warden 2000), as are<br />

10%<strong>–</strong>14% of biology majors (Moore and Cotner 2009). Answers<br />

in Genesis’ (AiG) Creation Museum, along with the $27 million in<br />

donations required to build it, attest to the appeal of young-Earth<br />

creationism. Indeed, AiG’s income for 2005 exceeded $13 million,<br />

and that of the Institute for Creation Research (ICR, another<br />

religious organization based on young-Earth creationism) exceeded<br />

$7 million. For comparison, the 2005 income of National<br />

Center for Science Education—the nation’s leading organization<br />

that defends the teaching of evolution in public schools—was<br />

C○ 2010 The Author(s). Journal compilation C○ 2010 The Society for the Study of Evolution.<br />

Evolution 64-3: 858<strong>–</strong>864


$1.2 million (Moore et al. 2009a). Clearly, Earth’s age remains<br />

one of the “sorest troubles” for many people today, just as it did<br />

for Darwin.<br />

In this study, we examined how college students’ selfdescribed<br />

religious and political views influence their beliefs<br />

about Earth’s age and how this may affect their knowledge and<br />

acceptance of evolution. To our knowledge, this is the first study<br />

to examine these factors in college students.<br />

Methods<br />

POPULATION AND SURVEY INSTRUMENT<br />

In 2009, we electronically surveyed 400 students enrolled in several<br />

sections of a nonmajors introductory biology course at the<br />

University of Minnesota. Because the course is one of a few options<br />

required of all undergraduates at the University, we assumed<br />

the survey population is representative of all students (politically,<br />

religiously, and demographically), except for those enrolled in the<br />

College of Biological Sciences. The optional survey, which was<br />

completed before the start of classes, consisted of the 20-item<br />

Measure of Acceptance of the Theory of Evolution (MATE) developed<br />

and validated by Rutledge and Sadler (2007), our own 10item<br />

Knowledge of Evolution Exam (KEE; Moore et al. 2009a),<br />

and several items intended to gauge students’ religious and political<br />

preferences. The MATE consists of statements such as “the age<br />

of the Earth is less than 20,000 years” and “humans are the product<br />

of evolutionary processes,” to which students noted their level of<br />

agreement on a five-point Likert scale (from “strongly agree” to<br />

“strongly disagree”). The KEE questions were modified from an<br />

internal exam database, were authentic to the nonmajors’ test experience,<br />

and were designed to evaluate students’ understanding<br />

of basic tenets of evolutionary theory—for example, how fitness<br />

is measured, natural selection as a mechanism for evolutionary<br />

change, etc. (Appendix I). None of the KEE items specifically addresses<br />

the age of the Earth or evolutionary time. We also asked<br />

students whether they were politically liberal, conservative, or<br />

middle-of-the-road, and whether they were not religious (atheist<br />

or agnostic), or, if religious, were they progressive/liberal, orthodox,<br />

or middle-of-the-road. Students could ignore questions, and<br />

their responses had no impact on students’ grades. Response rate<br />

varied by survey item, with as many as 195 students (almost 50%<br />

of the targeted group) completing the KEE, and as few as 124 responding<br />

to some of the MATE items. The survey and subsequent<br />

data collection were approved by the University of Minnesota’s<br />

Institutional Review Board.<br />

DATA ANALYSIS<br />

We extracted six variables from the survey to explore the factors<br />

that contribute to holding young-Earth and old-Earth beliefs about<br />

the origins of the world, on the one hand, and the relation of those<br />

OUTLOOK ON EVOLUTION AND SOCIETY<br />

beliefs to students’ knowledge of evolution and their presidential<br />

vote.<br />

The first two variables are self-reported measures of religious<br />

and political beliefs. The variable measuring students’ religious<br />

views (MYRELVIEW) is based on the question, “In general, I<br />

would describe my religious views as conservative, middle-ofthe-road,<br />

liberal/progressive, or none of the above/I’m not religious.”<br />

Responses were coded on a four-point scale from conservative<br />

(1) to not religious (4). Similarly, students’ political views<br />

(MYPOLVIEW) were determined by the question, “In general,<br />

I would describe my political views as conservative, middle-ofthe-road,<br />

or liberal.” MYPOLVIEW consists of a three-point scale<br />

ranging from conservative (1) to liberal (3).<br />

We constructed the young-Earth and old-Earth variables from<br />

MATE items that represent perspectives aligning with those beliefs.<br />

For the young-Earth variable (YOUNGEARTH), we averaged<br />

responses to the following two items: “The age of the Earth<br />

is less than 20,000 years,” and “The theory of evolution cannot be<br />

correct because it disagrees with the Biblical account of creation.”<br />

The old-Earth variable (OLDEARTH) was constructed by averaging<br />

responses to how much students agreed or disagreed with the<br />

statements that “Organisms existing today are the result of evolutionary<br />

processes that have occurred over millions of years,”<br />

and “The age of the earth is at least 4 billion years.” Scales for<br />

YOUNGEARTH and OLDEARTH range from 1 (Strongly Disagree)<br />

to 4 (Strongly Agree). Factor analysis confirms that each<br />

pair of items load onto a single dimension representing young-<br />

Earth creationist and old-Earth evolutionist beliefs with rotated<br />

factor loadings of 0.6979 and 0.7787, respectively.<br />

Students’ recollections of their high school biology courses<br />

(HSBIO) were captured in responses to an item asking them to<br />

identify whether or not evolution or creationism was taught in<br />

their courses. For the purposes of our analysis, HSBIO is coded<br />

according the following scheme: included neither evolution nor<br />

creationism = 1; included creationism, but not evolution = 2;<br />

included both evolution and creationism = 3; and included evolution,<br />

but not creationism = 4.<br />

The variable measuring students’ level of evolutionary biology<br />

knowledge (EVOGRADE) is a summative index of the<br />

number of questions answered correctly about various facets of<br />

evolutionary theory. EVOGRADE ranges in value from zero to 10<br />

with each increment representing a correctly answered question.<br />

The 2008 presidential candidate supported by each student was<br />

collected via the self-reported, retrospective response to the statement,<br />

“In the past presidential election, I voted/would have voted<br />

for John McCain or Barack Obama.” For analytical purposes, we<br />

constructed the dichotomous variable VOTE in which support for<br />

John McCain = 0 and support for Barack Obama = 1. Students<br />

who voted for other candidates constituted a small minority and<br />

were excluded from analysis.<br />

EVOLUTION MARCH 2010 859


OUTLOOK ON EVOLUTION AND SOCIETY<br />

Figure 1. Structural equation model demonstrating the nature of the relationships among the variables identified in Table 1. Note that<br />

“+” and“−” refer to positive and negative relationships, respectively, and asterisks follow the code ∗P < 0.05, ∗∗P< 0.01, ∗∗∗P < 0.001.<br />

Although our initial goals included the construction of a<br />

structural equation model (SEM) that employed simultaneous<br />

equations, the nature of the data is prohibitive (none of our variables<br />

is either linear or continuous). Therefore, we employed<br />

several individual models, each of which is best suited to the<br />

particular dependent variable under consideration. We then developed<br />

a structural model that demonstrates the nature of the<br />

relationships between and among our variables (Fig. 1). With<br />

the exception of models with VOTE as the dependent variable,<br />

all models are ordered logistical regression. We employed<br />

logistic regression models when the dichotomous VOTE variable<br />

was dependent. A Monte Carlo simulation was used to<br />

transform the ordered logistical regression results into predicted<br />

probabilities.<br />

Summary statistics for the variables under consideration are<br />

reported in Table 1.<br />

Table 1. Summary statistics.<br />

Variable N Mean Std. Dev. Min. Max.<br />

Myrelview 180 2.8111 0.9501 1 4<br />

Mypolview <strong>17</strong>3 2.3468 0.7361 1 3<br />

Youngearth 124 1.4839 0.6806 1 4<br />

Oldearth 132 3.3674 0.6958 1 4<br />

Hsbio 194 3.3402 1.0270 1 4<br />

Evograde 195 5.3026 2.2328 0 10<br />

Vote <strong>17</strong>4 0.7816 0.4143 0 1<br />

860 EVOLUTION MARCH 2010<br />

Results<br />

The first set of relationships includes the exogenous variables<br />

related to students’ religious and political views. These variables<br />

mutually reinforce one another at statistically significant levels<br />

(P < 0.001). That is, the more liberal one’s political views, the<br />

more likely one is to be liberal, agnostic, or atheistic in their<br />

religious views and vice versa.<br />

Students’ religious views also served as significant predictors<br />

of their beliefs about the origins of the world, their knowledge<br />

of evolutionary theory, and for whom they cast their presidential<br />

votes. Specifically, the more conservative a student’s religious<br />

views, the greater the likelihood of endorsing young-Earth beliefs<br />

(P < 0.05) and the less likely they are to endorse old-Earth<br />

evolutionist beliefs (P < 0.01). Additionally, more liberal, agnostic,<br />

or atheistic religious students were significantly more likely<br />

(P < 0.05) to correctly answer knowledge-based questions about<br />

theories and facts related to evolution. Finally, students holding<br />

less conservative religious views were considerably more likely<br />

to have voted for or supported Barack Obama in the 2008 presidential<br />

election (P < 0.001).<br />

The political views held by students contribute significantly<br />

to their disposition toward evolutionary theory and to their choice<br />

of presidential candidate. Here, the more liberal a student’s political<br />

views, the more likely they are to hold old-Earth beliefs about<br />

the origins of the world (P < 0.05) and to have cast their vote<br />

for Obama (P < 0.001). However, although more liberal political<br />

views are negatively related to acceptance of young-Earth views


and positively related to knowledge of evolutionary theory, political<br />

views fail to predict significantly either of those variables.<br />

Turning to the intermediate variables measuring the impact<br />

of beliefs about the origins of life on Earth, we find mutually<br />

reinforced inverse relationships between young-Earth and old-<br />

Earth views that are highly significant (P < 0.001). Specifically,<br />

those who hold young-Earth views are significantly less likely to<br />

accept an old-Earth rooted in evolutionary theory and vice versa.<br />

We also find that holding old-Earth beliefs contributes significantly<br />

to the ability of students to comprehend complex theoretical<br />

and factual tenets of evolutionary theory (P < 0.05).<br />

Furthermore, acceptance of old-Earth beliefs also significantly<br />

increased the probability of voting for or supporting Obama.<br />

Conversely, holding young-Earth beliefs appears to impact<br />

negatively, but not significantly, one’s ability to grasp cognitively<br />

evolutionary theory. And while holding creationist beliefs did<br />

predict a greater likelihood of voting for John McCain, they did<br />

not do so in a statistically significant manner.<br />

Finally, in keeping with previous findings (Moore and Cotner<br />

2009), the content of students’ high school biology courses<br />

was a significant predictor of their acceptance of evolutionary theory.<br />

The content of students’ high school biology course affects<br />

knowledge of evolution as well: students whose high school biology<br />

course included only evolutionary theory had approximately<br />

a 70% chance of answering half of the questions or more correctly<br />

whereas those with courses teaching only creationism had an approximately<br />

50% chance of doing so (for discussion, see Moore<br />

et al. 2009b); those with neither evolution nor creationism had a<br />

42% chance of scoring 50% or above, suggesting that creationismonly<br />

education is comparable to no education at all with respect to<br />

evolutionary biology. In combination, students who recall being<br />

taught evolution only in high school, and who are religiously and<br />

politically liberal, were more likely to earn any score above 50%<br />

on the exam than were their counterparts with more conservative<br />

educational, religious, and political backgrounds.<br />

Discussion<br />

College students have a variety of religious and political views that<br />

have been shaped collectively by their families, their communities,<br />

and the institutions with which they have come into contact<br />

during the course of their lives. These deeply rooted views influence<br />

students’ receptiveness to theories about life’s origins.<br />

And because these beliefs about creationism and evolution are<br />

firmly grounded in, and are expressions of, their worldviews, as<br />

instructors we are naïve to assume that students in college biology<br />

courses are necessarily open to scientific inquiry.<br />

Although student views of evolution are subject to multiple<br />

influences, our data indicate that their views about Earth’s age<br />

are a strong predictor of several different factors. But is the age<br />

OUTLOOK ON EVOLUTION AND SOCIETY<br />

of Earth one of our “sorest troubles?” Our research and historical<br />

and contemporary literature suggest the following:<br />

(1) Deep time is conceptually difficult to grasp, for the general<br />

public, science educators, and students throughout the educational<br />

spectrum.<br />

Charles Darwin himself had a hard time grasping “deep<br />

time”—the geologic concept of time, requiring billions of years—<br />

but knew that it was essential for his proposed evolutionary mechanism.<br />

Darwin didn’t publish any estimates of Earth’s age, but in<br />

the first edition of On the Origin of Species he did estimate the<br />

time needed to erode the Weald (a region in south England stretching<br />

from Kent to Surrey), “say three hundred million years.” He<br />

then explained: “I have made these few remarks because it is<br />

highly important for us to gain some notion, however, imperfect,<br />

of the lapse of years. During each of these years, over the whole<br />

world, the land and the water has been peopled by hosts of living<br />

forms. What infinite number of generations, which the mind<br />

cannot grasp, must have succeeded each other in the long roll of<br />

years” (Darwin 1859).<br />

More recently, attention has focused on the inability of<br />

scientists (Brush 2001), science educators (Petcovic and Ruhf<br />

2008), college students (Catley and Novick 2009; Libarkin 2006;<br />

Truscott et al. 2006), and the general public (Hofstadter 1996)<br />

to comprehend time in billions of years. Our deficiencies involve<br />

concepts about both absolute time (time in years; Catley and<br />

Novick 2009) and relative time (events in an accurate sequence;<br />

Libarkin 2006).<br />

(2) Students’ inability to accept an old Earth is a barrier to<br />

evolution acceptance.<br />

Not only is an appreciation of deep time foundational to<br />

our full understanding of life’s origin and diversification (Catley<br />

and Novick 2009; Hillis 2007), it is a critical concept for understanding<br />

geology, physics, and astrophysics (see review in Trend<br />

2002). Creationists themselves acknowledge that “An old age for<br />

the Earth is the heart of evolution” (Henry 2003), and have, in<br />

recent years, focused their argument to undermine an ancient<br />

Earth (Humphreys 1999; Baumgardner 2003; Humphreys et al.<br />

2004; DeYoung et al. 2005). Investing energy in debunking an<br />

ancient Earth is only logical if in fact Earth’s age is key to the<br />

Darwinian revolution, and is, in the words of Ernst Mayr (1972),<br />

“the snowball that started the whole avalanche.”<br />

Furthermore, evolution instruction is often independent of<br />

attempts to teach about deep time (Libarkin et al. 2005), a practice<br />

that enables students to learn about evolution without fully<br />

realizing its value as foundational to modern science. For example,<br />

Libarkin et al. (2005) describe students who can cite Earth’s<br />

age correctly but fail to comprehend the span of time Earth was<br />

uninhabitable, or the types of primitive organisms that likely arose<br />

first. These disconnects may speak to a need to incorporate historical<br />

arguments (Jensen and Finley 1995) or, more generally, an<br />

EVOLUTION MARCH 2010 861


OUTLOOK ON EVOLUTION AND SOCIETY<br />

emphasis on the nature of science (Lombrozo et al. 2008) into a<br />

more multidisciplinary strategy for teaching about evolution and<br />

the age of Earth.<br />

(3) Creationists’ explanations for life’s origin are easier to<br />

teach, learn and internalize than are scientific explanations that<br />

rely on an understanding of deep time.<br />

For evidence of creationism’s staying power, one can look to<br />

the recycled “argument for design” throughout history: In 45 BCE,<br />

Marcus Tullius Cicero invoked a sundial to claim evidence of a<br />

creator; in 1691, Anglican clergyman John Ray upgraded to a<br />

clock; and in 1802, William Paley’s treatise “Natural Theology”<br />

invoked a watch and its associated watchmaker (“every indication<br />

of contrivance, every manifestation of design, which existed in<br />

the watch, exists in the works of nature ...”). Technology<strong>–</strong>and<br />

the scientific advancements required to progress from sundials<br />

to watches—has changed, but the basic argument—evidence of<br />

design—remains unchanged (Moore et al. 2009a).<br />

Public-opinion polls provide more contemporary evidence<br />

of the power of creationists’ claims. In 1982, Philip Abelson<br />

wrote that scientific creationists “have no substantial body of<br />

experimental data to back their prejudices. Truth is not on their<br />

side. In the end, their activities must bring only harm to their<br />

cause.” At that time, 44% of the American public believed that<br />

living organisms were created in their present form within the<br />

last 10,000 years (Gallup 1982). After more than two decades of<br />

science education reform, that percentage remains unchanged.<br />

Whether we are constrained by teleology or intentionality<br />

(Sinatra et al. 2008), humans tend to believe in a universe designed<br />

for a purpose (Kelemen 2004; Evans 2001). This tendency<br />

lends a natural authority to supernatural explanations, which, in<br />

combination with a more intuitive model of Earth’s age (thousands,<br />

rather than billions, of years; see DeYoung et al. 2005),<br />

severely jeopardizes a realization of deep time. In asking our students<br />

to forsake the young, the intended, and the concrete for the<br />

old, the accidental, and the abstract, we may be fighting our own<br />

evolutionary history.<br />

(4) Teaching about time requires teaching for conceptual<br />

change.<br />

Learning about deep time is a conceptual change for most<br />

students, whereby instructors must realize the real barrier that<br />

Earth’s age presents to student understanding of evolution. In<br />

addition, and in accordance with studies of teaching for conceptual<br />

change (e.g., Sinatra et al. 2008), teachers may need to guide<br />

students to a realization that their understandings of Earth’s age,<br />

both absolute and relative, are faulty and in need of revision.<br />

Although several authors have suggested strategies for addressing<br />

relative time (e.g., James and Clark’s “ticking toilet paper roll”<br />

[2006]), input on conquering the absolute-time barrier is scarce.<br />

(5) Failure to accept an ancient Earth has real-world implications.<br />

862 EVOLUTION MARCH 2010<br />

We do not doubt that adherence to creationist explanations<br />

for natural phenomena has far-reaching significance. However,<br />

in this study, we specifically explored available information on<br />

voting activity in November 2008, and demonstrate a significant<br />

link between presidential candidate choice and Earth’s-age<br />

dimensions. This association is likely bolstered by contemporary<br />

partisan politics; for example, Republican party platforms<br />

in several states support some form of teaching creationism (as<br />

“Creation Science,” “Intelligent Design,” etc.) as a viable alternative<br />

theory to evolution by natural selection (e.g., Governor Sarah<br />

Palin’s Alaska: “We support giving Creation Science equal representation<br />

with other theories of the origin of life. If evolution is<br />

taught, it should be presented as only a theory”).<br />

The political motivation for embracing creationism is clear:<br />

Recent Gallup polls highlight the different levels of acceptance<br />

of evolution among Republicans, Democrats, and Independents<br />

(Newport 2008, 2009). In one poll, 60% of Republicans claim that<br />

humans were created in their present form by God within the last<br />

10,000 years, a belief shared by 40% of Independents and 38% of<br />

Democrats (Gallup 2008). Of those polled agreeing that humans<br />

evolved and God had no part, 4% are Republicans, 19% are Independents,<br />

and <strong>17</strong>% are Democrats. These data are consistent with<br />

findings that Republicans are more likely to attend church regularly,<br />

and Americans who attend church weekly are highly likely<br />

to believe in creationist explanations for human origins (Newport<br />

2006, 2008). For example, of those who attend church weekly,<br />

70% believe that God created humans in their present form (a<br />

central tenet of young-Earth creationists, or YEC); of those who<br />

seldom or never attend church, 24% are YEC (Newport 2008).<br />

These data suggest that religiosity correlates with one’s tendency<br />

to vote Republican and one’s likelihood to doubt evolutionary<br />

interpretations of origins, specifically, as highlighted herein,<br />

an Earth that is 4<strong>–</strong>5 billion years old. We do not intend to claim that<br />

ignorance of science is restricted to Republicans, nor that there<br />

are no creationist Democrats; rather, Republicans frequently embrace<br />

creationism more explicitly than do their counterparts and<br />

have benefited in recent years by making creationism a campaign<br />

issue.<br />

Implications<br />

Our research suggests that students who are liberal, agnostic,<br />

or atheistic in their religious views, are politically liberal, were<br />

taught evolution in high school, and accept the science behind<br />

evolutionary theory are more likely to understand the theoretical<br />

concepts and empirical findings related to evolution than those<br />

who are more conservative politically and religiously, received<br />

either no evolution or a diluted form of evolution instruction in<br />

high school, and who do not accept an old Earth. However, our<br />

findings also reveal a possible exception to this latter formulation:<br />

holding young-Earth views may not significantly impede a


student’s ability to learn facts about evolutionary theory. Thus,<br />

although it is not the role of biology instructors to engage in<br />

political or religious proselytizing, there remains the possibility<br />

of changing what students know about evolution via academic<br />

instruction.<br />

A student’s vote is likely a proxy for religious views, and<br />

therefore the association between their voting and their acceptance<br />

of evolution appears robust and bears further investigation.<br />

Given that adherence to young-Earth beliefs requires a refutation<br />

not only of modern biology, but also geology, paleontology,<br />

and physics, such convictions may serve as a proxy for scientific<br />

ignorance in general. Consequently, a commitment to comprehensive<br />

conceptual-change instruction in Earth’s age, the scientific<br />

method, and evolutionary theory could have practical, real-world<br />

implications that include nothing less than who we elect for political<br />

office.<br />

LITERATURE CITED<br />

Abelson, P. H. 1982. Creationism and the age of the earth. Science 215:119.<br />

Baumgardner, J. R. 2003. Catastrophic plate tectonics: the physics behind<br />

the genesis flood. Pp. 1<strong>–</strong>10 in In Proceedings of the Fifth <strong>International</strong><br />

Conference on Creationism, Pittsbugh, PA. Creation Science Fellowship,<br />

Inc. Pittsburgh, PA.<br />

Berkman, M. B., J. S. Pacheco, and E. Plutzer. 2008. Evolution and creationism<br />

in america’s classrooms: a national portrait. Plos Biol. 6:e124.<br />

Brush, S. G. 2001. Is the earth too old? The impact of geochronology on<br />

cosmology, 1929<strong>–</strong>1952. Geol. Soc. Lond. Spec. Publ. 190:157<strong>–</strong><strong>17</strong>5.<br />

Catley, K. M., and L. R. Novick. 2009. Digging deep: exploring college students’<br />

knowledge of macroevolutionary time. J. Res. Sci. Teach. 46:311<strong>–</strong><br />

332.<br />

Darwin, C. 1859. The origin of species. John Murray Publishers, London.<br />

DeYoung, D. B., J. Baumgardner, D. R. Humphreys, A. Snelling, S. A. Austin,<br />

E. F. Chaffin, S. W. Boyd, and L. Vardiman. 2005. Thousands, not<br />

billions: challenging an icon of evolution: questioning the age of the<br />

Earth. New Leaf Publishing Group, Green Forest, AR.<br />

Evans, E. M. 2001. Cognitive and contextual factors in the emergence of diverse<br />

belief systems: creation versus evolution. Cogn. Psychol. 42:2<strong>17</strong><strong>–</strong><br />

266.<br />

Gallup, Poll 1982. Evolution, Creationism, Intelligent Design. Available at:<br />

http://www.gallup.com/poll/21814/Evolution-Creationism-Intelligent-<br />

Design.aspx. Accessed December 7, 2009.<br />

Henry, J. 2003. An old age for the earth is the heart of evolution. Creation<br />

Res. Soc. Q. 40:164<strong>–</strong><strong>17</strong>2.<br />

Hillis, D. M. 2007. Making evolution relevant and exciting to biology students.<br />

Evolution 61:1261<strong>–</strong>1264.<br />

Hofstadter, D. R. 1996. Metamagical themas: questing for the essence of mind<br />

and pattern. illustrated ed. Basic Books. Web, New York.<br />

Humphreys, D. R. 1999. Evidence for a young world. Creation Matters 4:1<strong>–</strong>4.<br />

Humphreys, D. R., S. A. Austin, J. R. Baumgardner, and A. A. Snelling. 2004.<br />

Helium diffusion age of 6,000 years supports accelerated nuclear decay.<br />

Creation Res. Soc. Q. 41:1<strong>–</strong>16.<br />

James, P., and I. Clark. 2006. Overcoming geological misconceptions. Planet<br />

<strong>17</strong>:10<strong>–</strong>13.<br />

Jensen, M. S., and F. N. Finley. 1995. Teaching evolution using historical<br />

arguments in a conceptual change strategy. Sci. Educ. 79:147<strong>–</strong>167.<br />

Kelemen, D. 2004. Are children “intuitive theists”? Reasoning about purpose<br />

and design in nature. Psychol. Sci. 15:295<strong>–</strong>301.<br />

OUTLOOK ON EVOLUTION AND SOCIETY<br />

Libarkin, J. C. 2006. College student conceptions of geological phenomena<br />

and their importance in classroom instruction. Planet <strong>17</strong>:6<strong>–</strong>9.<br />

Libarkin, J. C., S. W. Anderson, J. Dahl, M. Beilfuss, W. Boone, and J. P.<br />

Kurdziel. 2005. Qualitative analysis of college students’ ideas about<br />

the earth: interviews and open-ended questionnaires. J. Geosci. Educ.<br />

53:<strong>17</strong><strong>–</strong>26.<br />

Lombrozo, T., A. Thanukos, and M. Weisberg. 2008. The importance of<br />

understanding the nature of science for accepting evolution. Evol. Educ.<br />

Outreach 1:290<strong>–</strong>298.<br />

Mayr, E. 1972. The nature of the Darwinian revolution. Science <strong>17</strong>6:981<strong>–</strong>989.<br />

Moore, R., and S. Cotner. 2009. The creationist down the hall: does it matter<br />

when teachers teach creationism? BioScience 59:429<strong>–</strong>435.<br />

Moore, R., S. Cotner, and A. Bates. 2009a. The influence of religion and high<br />

school biology courses on students’ knowledge of evolution when they<br />

enter college. J. Excellent Teach. (Special Issue on Evolution Education),<br />

9:3<strong>–</strong>11.<br />

Moore, R., M. D. Decker, and S. Cotner. 2009b. No prospect of an end: a<br />

chronology of the evolution-creationism controversy. Greenwood Press,<br />

Westport, CT.<br />

Newport, F. 2006. Almost half of Americans believe humans did not<br />

evolve. Available at: http://www.gallup.com/poll/23200/Almost-Half-<br />

Americans-Believe-Humans-Did-Evolve.aspx. Accessed December 7,<br />

2009.<br />

———. 2008. Republicans, democrats differ on creationism. Available<br />

at: http://www.gallup.com/poll/108226/Republicans-Democrats-Differ-<br />

Creationism.aspx. Accessed December 7, 2009.<br />

———. 2009. On Darwin’s birthday, only 4 in 10 believe in evolution.<br />

Retrieved May 3, 2009, from http://www.gallup.com/poll/<br />

114544/Darwin_Believe-Evolution.aspx<br />

Petcovic, H. L., and R. J. Ruhf. 2008. Geoscience conceptual knowledge<br />

of preservice elementary teachers: results from the geoscience concept<br />

inventory. J. Geosci. Educ. 56:251<strong>–</strong>260.<br />

Rutledge, M. L., and M. A. Warden. 2000. Evolutionary theory, the nature<br />

of science and high school biology teachers: critical relationships. Am.<br />

Biol. Teacher 62:23<strong>–</strong>31.<br />

Rutledge, M. L., and K. C. Sadler. 2007. Reliability of the Measure of Acceptance<br />

of the Theory of Evolution (MATE) instrument with university<br />

students. Am. Biol. Teacher 69:332<strong>–</strong>335.<br />

Sinatra, G. M., S. K. Brem, and E. M. Evans. 2008. Changing minds? Implications<br />

of conceptual change for teaching and learning about biological<br />

evolution. Evol. Educ. Outreach 1:189<strong>–</strong>195.<br />

Trend, R. D. 2002. Developing the concept of deep time. Pp. 187<strong>–</strong>201 in V.<br />

J. Mayer, ed. Global science literacy. Springer, New York.<br />

Truscott, J. B., A. Boyle, S. Burkill, J. Libarkin, and J. Lonsdale. 2006. The<br />

concept of time: can it be fully realised and taught? Planet <strong>17</strong>:21<strong>–</strong>23.<br />

Associate Editor: T. Meagher<br />

Appendix I<br />

KNOWLEDGE OF EVOLUTION EXAM (KEE)<br />

(1) Which of the following support the theory of evolution?<br />

A. artificial selection (also known as selective breeding), an<br />

analogue of natural selection<br />

B. comparative biochemistry, where similarities and differences<br />

of DNA among species can be quantified<br />

C. vestigial structures that serve no apparent purpose<br />

D. comparative embryology, where the evolutionary history<br />

of similar structures can often be traced<br />

EVOLUTION MARCH 2010 863


OUTLOOK ON EVOLUTION AND SOCIETY<br />

E. all of the above provide evidence to support the theory of<br />

evolution<br />

(2) Resistance to a wide variety of insecticides has recently<br />

evolved in many species of insects. Why?<br />

A. mutations are on the rise<br />

B. humans are altering the environments of these organisms,<br />

and the organisms are evolving by natural selection<br />

C. no new species are evolving, just resistant strains or varieties.<br />

This is not evolution by natural selection<br />

D. humans have better health practices, so these organisms<br />

are trying to keep up<br />

E. insects are smarter than humans<br />

(3) Which of the following is the most fit in an evolutionary sense?<br />

A. a lion who is successful at capturing prey but has no cubs<br />

B. a lion who has many cubs, eight of which live to adulthood<br />

C. a lion who overcomes a disease and lives to have three<br />

cubs<br />

D. a lion who cares for his cubs, two of whom live to adulthood<br />

E. a lion who has a harem of many lionesses and one cub<br />

(4) How might a biologist explain why a species of birds has<br />

evolved a larger beak size?<br />

A. large beak size occurred as a result of mutation in each<br />

member of the population<br />

B. the ancestors of this bird species encountered a tree with<br />

larger than average sized seeds. They needed to develop<br />

larger beaks to eat the larger seeds, and over time, they<br />

adapted to meet this need<br />

C. some members of the ancestral population had larger<br />

beaks than others. If larger beak size was advantageous,<br />

they would be more likely to survive and reproduce. As<br />

such, large beaked birds increased in frequency relative<br />

to small beaked birds<br />

D. the ancestors of this bird species encountered a tree with<br />

larger than average sized seeds. They discovered that by<br />

stretching their beaks, the beaks would get longer, and<br />

this increase was passed on to their offspring. Over time,<br />

the bird beaks became larger<br />

E. none of the above<br />

(5) Which of the following statements about natural selection is<br />

true?<br />

A. natural selection causes variation to arise within a population<br />

B. natural selection leads to increase likelihood of survival<br />

for certain individuals based on variation. The variation<br />

comes from outside the population<br />

C. all individuals within a population have an equal chance<br />

of survival and reproduction. Survival is based on choice<br />

864 EVOLUTION MARCH 2010<br />

D. natural selection results in those individuals within a population<br />

who are best-adapted surviving and producing<br />

more offspring<br />

E. natural selection leads to extinction<br />

(6) All organisms share the same genetic code. This commonality<br />

is evidence that<br />

A. evolution is occurring now<br />

B. convergent evolution has occurred<br />

C. evolution occurs gradually<br />

D. all organisms are descended from a common ancestor<br />

E. life began millions of years ago<br />

(7) Which of the following statements regarding evolution by<br />

natural selection is FALSE?<br />

A. natural selection acts on individuals<br />

B. natural selection is a random process<br />

C. very small selective advantages can produce large effects<br />

through time<br />

D. natural selection can result in the elimination of certain<br />

alleles from a population’s gene pool<br />

E. mutations are important as the ultimate source of genetic<br />

variability upon which natural selection can act<br />

(8) A change in the genetic makeup of a population of organisms<br />

through time is<br />

A. adaptive radiation<br />

B. biological evolution<br />

C. LaMarckian evolution<br />

D. natural selection<br />

E. genetic recombination<br />

(9) Which of the following is the ultimate source of new variation<br />

in natural populations?<br />

A. recombination<br />

B. mutation<br />

C. hybridization<br />

D. gene flow<br />

E. natural selection<br />

(10) Which of the following best describes the relationship between<br />

evolution and natural selection?<br />

A. natural selection is one mechanism that can result in the<br />

process of evolution<br />

B. natural selection produces small-scale changes in populations,<br />

whereas evolution produces large-scale ones<br />

C. natural selection is a random process whereas evolution<br />

proceeds toward a specific goal<br />

D. natural selection is differential survival of populations or<br />

groups, resulting in the evolution of individual organisms<br />

E. they are equivalent terms describing the same process


Conservation and<br />

Behaviour<br />

Richard Buchholz, Department of Biology, University of Mississippi, Mississippi, USA<br />

Patrick Yamnik, Department of Biology, University of Mississippi, Mississippi, USA<br />

Cassan N Pulaski, Department of Biology, University of Mississippi, Mississippi, USA<br />

Chelsea A Campbell, Department of Biology, University of Mississippi, Mississippi, USA<br />

Conservation behaviour is an area of conservation biology particularly suited to<br />

investigate problems of species endangerment associated with managing animals in<br />

fragmented habitats and isolated parks. It employs a theoretical framework that<br />

examines the mechanisms, development, function and phylogeny of behavioural<br />

variation in order to develop practical tools for preventing extinction. This framework<br />

can be used to attract animals to suitable habitat, restore migratory movements,<br />

identify individual reproductive strategies affecting conservation and focus protection<br />

on species most susceptible to extinction. The future success of conservation behaviour<br />

requires that behaviourists link individual variation in behaviour to processes that<br />

determine population viability.<br />

Introduction<br />

Conservation behaviour is an emerging discipline that utilizes<br />

the framework and methods of ethology to help solve<br />

issues important to the biodiversity crisis (Buchholz, 2007).<br />

Species are becoming extinct or threatened with extinction<br />

at an accelerating rate due to human population growth<br />

and our unsustainable use of natural resources. The relatively<br />

new field of conservation behaviour aims to use a<br />

scientific understanding of animal behaviour to both determine<br />

the reasons for differential susceptibility to extinction,<br />

and to manage threatened animals and their habitats<br />

to promote population viability. Conservation behaviour<br />

is not merely the description of the natural history of an<br />

animal. Instead conservation behaviour uses the framework<br />

of ethology, originally posed by the Nobel laureate<br />

Niko Tinbergen, to allow conservation issues to be investigated<br />

from four complementary behavioural perspectives:<br />

mechanisms, ontogeny, adaptive function and<br />

phylogeny. The proximate mechanisms of behaviour include<br />

the physiological and neuro-endocrine underpinnings<br />

of animal action. The ontogeny of behaviour<br />

ELS subject area: Ecology<br />

How to cite:<br />

Buchholz, Richard; Yamnik, Patrick; Pulaski, Cassan N; and, Campbell,<br />

Chelsea A (December 2008) Conservation and Behaviour. In:<br />

Encyclopedia of Life Sciences (ELS). John Wiley & Sons, Ltd: Chichester.<br />

DOI: 10.1002/9780470015902.a00212<strong>17</strong><br />

. Introduction<br />

. Habitat Selection<br />

Advanced article<br />

Article Contents<br />

. Dispersal, Migration and Conservation<br />

. Reproductive Behaviours and Population Growth<br />

. Effects of Small Population Size on Behaviour<br />

. Managing Human<strong>–</strong>Wildlife Conflict<br />

. Implications for Management<br />

Online posting date: 15 th December 2008<br />

explores the genetic and learned basis of behavioural development<br />

over the lifetime of individual animals. Ultimately<br />

behaviour patterns may cause differential survival<br />

or reproduction of individuals, and thus serve an adaptive<br />

function. Finally the ultimate origins of inter-specific variation<br />

in the behaviour may be investigated by exploring<br />

how behaviour has evolved among a set of phylogenetically<br />

related species. These four approaches can be integrated to<br />

allow conservation behaviourists to propose novel solutions<br />

to conservation problems. See also: Biodiversity <strong>–</strong><br />

Threats; Modern Extinction; Tinbergen, Nikolaas<br />

Although the interdisciplinary field of conservation biology<br />

was formalized as a science in the mid-1980s with the<br />

formation of the Society for Conservation Biology, recognition<br />

that behavioural biology could contribute uniquely<br />

to conservation did not coalesce for another decade, as<br />

evidenced by a flurry of symposia and publications on the<br />

subject (listed by Caro, 2007). Conservation behaviour<br />

came about as there was an increasing realization that<br />

population-level investigations of genetics and ecology<br />

were insufficient to explain some rather dramatic species<br />

declines. For example, cheetah Acinonyx jubatus population<br />

decline was linked to cheetah-cub killing by lions<br />

Panthera leo and other large carnivores (Laurenson et al.,<br />

1995). Because cheetah-cub killing is not simply a form of<br />

predation for food, a more complex behavioural investigation<br />

of the origins and causes of this behaviour by lions is<br />

warranted. A conservation behaviour approach to this<br />

problem would include an investigation of the sensory cues<br />

that lions use to find cheetah litters (mechanism), the ability<br />

of cheetah mothers to learn to reduce their risk of litter<br />

predation (ontogeny), the effect of cheetah-cub killing on<br />

ENCYCLOPEDIA OF LIFE SCIENCES & 2008, John Wiley & Sons, Ltd. www.els.net 1


the lion’s survival and reproduction (adaptive function),<br />

and a comparative analysis of the ecological and life-history<br />

factors associated with inter-specific infanticide<br />

among species in the order Carnivora (phylogeny). This<br />

comprehensive approach to conservation holds promise<br />

for ameliorating conservation problems if it can be translated<br />

into management actions that help threatened populations<br />

grow to a sustainable equilibrium (Beissinger,<br />

1997).<br />

When a species is reduced to just a few animals, the behaviour<br />

of every individual has a direct and obvious impact<br />

on the species’ survival. When animals are endangered, but<br />

not at the brink of extinction, the role of behaviour may be<br />

less obvious but is no less important. Natural selection is<br />

expected to mould the decision-making of individual animals<br />

so that they weigh the potential lifetime fitness costs<br />

and benefits of their options before choosing a course of<br />

action. The conservation behaviourist’s objective is to understand<br />

how animal decisions may have become maladaptive<br />

in a human-dominated landscape, and to<br />

participate in management efforts that correct or compensate<br />

for behaviour that reduces survival or reproduction.<br />

Although there are myriad ways that behaviour might impact<br />

population dynamics, five behavioural topics will play<br />

major roles in protecting biodiversity: habitat selection,<br />

dispersal and migratory movements, reproduction, living<br />

in small populations and human<strong>–</strong>animal conflict over resource<br />

use. See also: Conservation of Populations and<br />

Species<br />

Habitat Selection<br />

Habitats that look fairly homogeneous to the untrained<br />

human eye may be heterogeneous in features important to<br />

the survival and reproduction of a particular animal species.<br />

Even within species, suitable habitat can vary with the<br />

age, sex and condition of the individual. Thus we expect<br />

animals to be choosy about where they live, and to move to<br />

places that maximize their lifetime reproductive success. If<br />

landscape changes by humans cause animals to bypass<br />

good habitats, or to become attracted to poor habitats,<br />

population decline may occur. The framework of conservation<br />

behaviour provides us with potential management<br />

tools to manipulate the habitat use of animals so that they<br />

use habitats optimally in light of anthropogenic alterations<br />

to which they have not had time to adapt in an evolutionary<br />

sense.<br />

Getting animals past human obstacles<br />

Goal-oriented movements of animals might cover very little<br />

geographic distance or alternatively occur for just a few<br />

minutes of the animal’s life, but be of great importance to<br />

the survival and reproductive success of individuals. For<br />

example, some species have a critically important transition<br />

from the location where they are born or hatched, and<br />

the place where they will be reared. Understanding the cues<br />

2<br />

Conservation and Behaviour<br />

that these naïve, fledgling animals use to find their way<br />

allows the conservationist to avoid conflicts between this<br />

critical transition and anthropogenic changes. For example,<br />

newly hatched sea turtles and fledgling petrels must<br />

move from their nests to the ocean, but can become disoriented<br />

by some wavelengths of artificial night lighting<br />

(Witherington, 1997). Likewise, roadways and hydroelectric<br />

dams can function as barriers to animal movement that<br />

could be circumvented through management of the cues<br />

that imperilled species use to find their way. There will be<br />

times when the findings of conservation behaviourists may<br />

run into conflict with other efforts at environmental protection.<br />

For example, migratory tree bats (Lasiurus species)<br />

appear to be using wind turbines as mating sites along migration<br />

routes. Cryan and Brown (2007) have suggested<br />

that these normally solitary bats may use a behavioural rule<br />

such as ‘go to the tallest structure’ to rendezvous during fall<br />

migration. As a result dozens or hundreds of bats are killed<br />

during a several week period, and the protection of biodiversity<br />

finds itself at odds with efforts to use alternative<br />

energy sources to reduce global climate change. Conservation<br />

behaviourists must be involved in efforts to make<br />

wind turbines less attractive or less deadly to bats, as well as<br />

participate in other situations where acute animal contact<br />

with human landscape structures cause mortality and interrupt<br />

crucial events in a species’ life cycle.<br />

Getting animals to use good habitat<br />

In human-dominated landscapes habitat patches are likely<br />

to be relatively small and susceptible to local extinctions.<br />

Small populations in those habitat fragments may decline<br />

for a number of reasons, including inbreeding depression<br />

and greater exposure to climatic extremes. Suitable patches<br />

in some cases may remain unoccupied if cues used in habitat<br />

selection are absent, if the patch is outside of an organism’s<br />

perceptual range or if misleading cues are present<br />

that deter settlement (Gilroy and Sutherland, 2007).<br />

Behavioural knowledge may be used to lure individuals<br />

to settle in these unoccupied or underutilized habitats.<br />

Many species display conspecific attraction wherein the<br />

presence of conspecifics is likely to be indicative of suitable<br />

habitat (Ahlering and Faaborg, 2006). Ward and<br />

Schlossberg (2004) successfully attracted endangered territorial<br />

black-capped vireos Vireo atricapilla to unoccupied<br />

habitat using recorded vocalizations. Heterospecific cues<br />

are also used in habitat selection; least flycatchers, Empidonax<br />

minimus were attracted to habitat utilized by<br />

American redstarts Setophaga ruticilla (Fletcher, 2007).<br />

Similarly, marbled newts Triturus marmoratus oriented towards<br />

the calls of natterjack toads Bufo calamita (Diego-<br />

Rasilla and Luengo, 2004). In addition to conspecific or<br />

heterospecific auditory cues, olfactory, chemical and visual<br />

cues may also be used to attract individuals to unoccupied<br />

habitat (Gilroy and Sutherland, 2007). Habitat cues such as<br />

light and vegetation are also used by organisms in selecting<br />

suitable habitat, and these cues may be manipulated to<br />

attract individuals to unoccupied patches. Similarly<br />

ENCYCLOPEDIA OF LIFE SCIENCES & 2008, John Wiley & Sons, Ltd. www.els.net


animals may adopt microhabitat choices by observing the<br />

behaviour of experienced individuals.<br />

Getting animals to avoid bad habitat<br />

Alternatively it may be necessary to dissuade animals from<br />

using an attractive habitat patch that is actually harmful.<br />

Such ‘ecological traps’ occur when cues that historically<br />

indicated suitable habitat have become disconnected from<br />

habitat quality due to human disturbance. An ecological<br />

trap functions as a population sink (Schlaepfer et al., 2002).<br />

The conservation behaviourist may develop tools for<br />

masking misleading cues to eliminate the sensory allure<br />

or introduce additional cues to indicate the habitat’s poor<br />

suitability for settlement. The majority of ecological trap<br />

studies have focused on birds, and a number of these have<br />

found forest edge habitats to function as traps due to increased<br />

predation and nest-parasitism (Battini, 2004).<br />

Edge-effect traps may be countered by creating openings<br />

in the canopy of the forest interior which may attract birds<br />

away from the habitat edge (Battini, 2004). Heterospecific<br />

cues may also be used to dissuade individuals from settling<br />

in unsuitable habitat. American redstarts S. ruticilla were<br />

more likely to avoid habitats with least flycatcher, E. minimus<br />

vocalizations than control habitat patches (Fletcher,<br />

2007). Understanding behavioural cues and mechanisms<br />

resulting in ecological traps will allow for the management<br />

of these problematic habitats and their associated wildlife.<br />

Dispersal, Migration and Conservation<br />

The direction, rate and motivational basis for animal<br />

movement varies throughout the day, seasonally and with<br />

life stage. If we employ the Tinbergen framework to the<br />

problem of an animal’s movement out of its traditional<br />

home range, conservation behaviour might consider, for<br />

example, the hormonal (mechanistic) basis for an individual’s<br />

decision to move elsewhere, whether it learned about<br />

the direction of movements from its parents (ontogeny)<br />

and the fitness consequences of the move (adaptive function).<br />

Comparing the relationship between life-history<br />

traits and the movement patterns of a group of related<br />

species might be useful in anticipating conservation conflicts.<br />

Individual differences in the distance that an animal<br />

moves provides additional structure for our discussion of<br />

how animal locomotion relates to conservation of biodiversity.<br />

Geographical scales of animal movement can be<br />

divided into two broad categories: dispersal over relatively<br />

short distances, and migration over much longer distances.<br />

Short distances<br />

Short distance dispersal entails emigration from the natal<br />

or home range wherein an individual’s objective for moving<br />

may be to avoid inbreeding, reduce competition with kin or<br />

to locate better food resources. There are a wide range of<br />

abiotic and biotic factors that determine dispersal. In<br />

fragmented or isolated habitat patches, dispersing individuals<br />

may suffer high mortality and reduce the viability of<br />

the local population. Dispersing animals, particularly large<br />

carnivores, may come into conflict with humans when they<br />

travel through human-dominated habitats, reducing public<br />

support for wildlife preservation. Thus investigating the<br />

motivation behind individual dispersal can be critical to<br />

species conservation. For example, Zedrosser et al. (2007)<br />

demonstrated that male brown bears Ursus arctos will migrate<br />

up to 119 km to avoid inbred mating, exposing them<br />

to a greater risk for human contact. Proper management of<br />

the brown bear might involve translocation of breedingage<br />

animals to create local breeding opportunities for bears<br />

that would otherwise disperse. Alternatively manipulating<br />

environmental cues to redirect the movement of dispersing<br />

bears through protected corridors may reduce bear mortality<br />

and conflict with humans. See also: Range Limits<br />

The causes of animal dispersal will be of increasing concern<br />

as disease epidemics and emerging zoonotic pathogens<br />

threaten both human and wildlife health. The social systems<br />

of chimpanzees Pan troglydytes and gorillas Gorilla<br />

gorilla, for example, may explain the differential impact of<br />

the ebola virus on these great ape species. In Great Britain<br />

Eurasian badgers Meles meles serve as reservoirs of bovine<br />

tuberculosis, and areas with high densities of badgers typically<br />

have higher incidences of bovine tuberculosis<br />

(McDonald et al., 2008). Reducing the density of badgers<br />

through culling, however, disrupts the social and territorial<br />

structure of the badgers. This resulted in the surviving<br />

badgers travelling greater distances and utilizing larger<br />

areas, effectively spreading bovine tuberculosis. The latter<br />

example highlights the need for a behaviourist’s understanding<br />

of individual behavioural strategies for anticipating<br />

the complexities of disease transmission in landscapes.<br />

Long distances<br />

Conservation and Behaviour<br />

Seasonal migration may involve non-stop movement<br />

across thousands of kilometres of unsuitable habitat, resulting<br />

in the mortality of individuals in poor condition or<br />

those unable to orient correctly. Unfortunately there is little<br />

information known about how the different stages of<br />

migratory movement affect survival and reproduction. For<br />

example, how might the migratory routes or duration of<br />

stopovers at resting sites affect fecundity in the subsequent<br />

breeding seasons? These migratory options are behavioural<br />

decisions suitable for investigation by conservation behaviourists.<br />

The most information available is for birds, with<br />

virtually no data available for the highly threatened migrations<br />

of hoofed mammals (Bolger et al., 2008). Comparative<br />

studies in birds have revealed that migratory<br />

species are more susceptible to decline or extinction than<br />

non-migratory species (references in Thomas et al., 2006).<br />

In a phylogenetically controlled analysis of population<br />

status in North American shorebirds, Thomas et al. (2006)<br />

found that species that opted for inland (rather than<br />

coastal) migratory routes were more threatened. The authors<br />

suggest that additional protections for ephemeral<br />

ENCYCLOPEDIA OF LIFE SCIENCES & 2008, John Wiley & Sons, Ltd. www.els.net 3


Number declining or extinct<br />

wetlands and upland area stopover sites might stem the<br />

decline of these bird species. Green sea turtles Chelonia<br />

mydas that both reproduce and forage in the Cocos Islands<br />

have very brief migrations compared to other populations,<br />

and are thus hypothesized to have spared energetic resources<br />

that should result in greater fecundity or survival<br />

(Whiting et al., 2008). If confirmed, this population of turtles<br />

might be more resilient to human disturbance, allowing<br />

conservationists to focus their preservation efforts on the<br />

more susceptible long-distance migrants.<br />

Migratory motivation and compass orientation can have<br />

both learned and instinctual components. New migratory<br />

routes seem to evolve rather quickly, and thus may not be<br />

reflected in broad brush genetic determinations of ‘evolutionary<br />

significant units’, a measure of population uniqueness<br />

favoured by policy makers (Davis et al., 2006). Efforts<br />

to recover the endangered whooping crane Grus americana<br />

in North America have entailed using behavioural biology<br />

to stop migration in some reintroduced populations (central<br />

Florida, USA), and establish new migratory routes<br />

(between north Florida and Wisconsin, USA) in others<br />

(Canadian Wildlife Service and U.S. Fish and Wildlife<br />

Service, 2005), in both cases through the manipulation of<br />

learned cues in the rearing environment. Virtually nothing<br />

is known about the mechanistic and developmental basis of<br />

migration in ungulates (Bolger et al., 2008); however, a<br />

better understanding of how their behaviour might be<br />

managed so that they can circumvent anthropogenic<br />

obstacles will be critical to the conservation of these land-bound<br />

mammals (Figure 1). See also: Migration: Orientation<br />

and Navigation<br />

Wildlife corridors<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Dispersal behaviour is of particular interest to conservation<br />

biologists due to its potential impacts on population<br />

Overhunting Habitat loss<br />

Threats<br />

Migration obstacle<br />

Figure 1 More migratory populations and species of hoofed mammals are endangered by anthropogenic obstacles to animal migration such as fences, roads<br />

and reservoirs, than by overhunting or agricultural development of natural areas. Data from Bolger et al. (2008).<br />

4<br />

Conservation and Behaviour<br />

genetics. Animals confined to small, disconnected populations<br />

have only genetic relatives to reproduce with, resulting<br />

in low genetic variability, inbreeding depression and<br />

poor disease resistance. Dispersal behaviour significantly<br />

affects population connectivity, and may be facilitated by<br />

the creation of suitable habitat corridors within the fragmented<br />

landscape. Behavioural studies can provide important<br />

insight into designing corridors based on<br />

behavioural responses of wildlife to habitat edges and corridors.<br />

See also: Landscape Ecology<br />

Behavioural studies will provide valuable insight into<br />

determining when corridors are needed and how to design<br />

them. Behavioural studies have demonstrated that many<br />

species, including spotted salamanders Ambystoma maculatum,<br />

and two butterfly species, Phoebis sennae and<br />

Eurema nicippe, detect and avoid habitat edges, and this<br />

behaviour leads to a greater probability of corridor use<br />

(Haddad, 1999; Rittenhouse and Semlitsch, 2006). The<br />

habitat quality within the corridor itself may determine its<br />

effectiveness. The Chucao tapaculo Scelorchilus rubecula<br />

disperses well through low-quality shrub habitat to access<br />

forest patches but avoids large open gaps between habitat<br />

patches (Castello´ n and Sieving, 2005). The salamander<br />

Ensatina eschscholtzii travels more rapidly between suitable<br />

habitat patches if corridor quality is poor (Rosenberg<br />

et al., 1998). Thus, behavioural studies can greatly inform<br />

the use and design of management by understanding which<br />

organisms are likely to use corridors effectively based on<br />

their response to habitat edges, their behaviour within corridors<br />

and other factors.<br />

An important application of corridor theory is the use of<br />

highway overpasses and underpasses to provide connectivity<br />

between habitat patches that are bisected by roads.<br />

Understanding how animals respond to roads, traffic and<br />

the structural design of under- and overpasses will allow<br />

managers to design corridors that maximize wildlife use<br />

ENCYCLOPEDIA OF LIFE SCIENCES & 2008, John Wiley & Sons, Ltd. www.els.net


and minimize traffic-induced wildlife mortality. For example,<br />

highway crossings by elk Cervus elaphus in Arizona,<br />

USA, decreased with increasing traffic volume, although<br />

crossings were more likely to occur near important resources<br />

and depended on the season (Gagnon et al., 2007).<br />

Importantly Ng et al. (2004) found that the structural design<br />

of crossings can greatly affect their utility. Over- and<br />

underpass design may be the most important factor affecting<br />

which, if any, species are likely to use them, with some<br />

species such as grizzly bears, U. arctos preferring wide,<br />

short and tall crossings, whereas other species such as cougars<br />

Puma concolor prefer more constricted passages<br />

(Clevenger and Waltho, 2005; Mata et al., 2005). Understanding<br />

an organism’s behaviour will allow managers to<br />

most efficiently plan highway crossings that will suit wildlife<br />

needs.<br />

Reproductive Behaviours and<br />

Population Growth<br />

The rate and productivity with which animals breed determine<br />

in part whether their population can compensate for<br />

losses due to anthropogenic changes such as climate<br />

change, hunting and competition with invasive species.<br />

Differential mating success, or sexual selection, occurs<br />

through mate choice and intra-sexual competition, with<br />

myriad implications for conservation (Quader, 2005).<br />

See also: Sexual Selection<br />

Mating systems and genetic diversity<br />

Mating systems provide broadly defined categories encompassing<br />

how the sexual strategies of females and males<br />

are integrated under natural conditions. Polygamous mating<br />

systems have been linked to an increased risk of extinction<br />

in birds (Doherty et al., 2003), perhaps because the<br />

effective population size (N e) is lower in mating systems<br />

where only a few males are mating with the majority of<br />

females (Parker and Waite, 1997). In small populations in<br />

particular low genetic diversity may be exacerbated by extreme<br />

differences in male mating success, an effect that may<br />

be ameliorated by behavioural management. For example,<br />

Alberts et al. (2002) found that the temporary removal of<br />

dominant males in the threatened Cuban iguana (Cyclura<br />

nubile) allowed other males to mate, theoretically increasing<br />

genetic variability and population persistence. If female<br />

choice is the primary determinant of male mating success,<br />

however, it might be appropriate to manipulate positively<br />

the apparent quality of a male not normally attractive to<br />

females. Fisher et al. (2003) showed that by redistributing<br />

scent markings of individual male loris in captivity, it was<br />

possible to increase the likelihood that females would subsequently<br />

mate with them.<br />

Human efforts to promote conservation can have inadvertent<br />

negative effects on population viability when they<br />

focus on natural selection and ignore sexual selection.<br />

Trophy hunting has been traditionally thought of as a relatively<br />

benign way to both remove surplus males from<br />

populations and provide financial resources to human<br />

communities surrounding protected areas. Infanticide by<br />

males that replace trophy hunted males, unfortunately,<br />

may destabilize lion prides. Thus the intra-sexual competition<br />

of male lions enabled by trophy hunting may result in<br />

unsustainable use if male age is not considered (Whitman<br />

et al., 2004). In the case of bighorn sheep Ovis canadensis<br />

removal of males with trophy size horns removes fastgrowing<br />

males whose ‘good genes’ may normally benefit<br />

female reproductive success indirectly through the survival<br />

of their offspring (Coltman et al., 2003). Even when wildlife<br />

managers attempt to positively affect female condition, for<br />

example, through supplemental feeding, sexual selection<br />

can cause unexpected population consequences. Supplemental<br />

feeding of female endangered kakapo parrots Strigops<br />

habroptilus in New Zealand resulted in male-biased<br />

clutches of offpring (Clout et al., 2002). Because males must<br />

be large to compete with other males for matings, mothers<br />

apparently produce more sons if they have an abundance of<br />

food. See also: Natural Selection: Sex Ratio<br />

Parental behaviour<br />

Conservation and Behaviour<br />

Parental care of offspring must also be considered in managing<br />

endangered animal populations. In some species,<br />

adults may be present at nesting sites and breeding burrows,<br />

but are not contributing their genes to the next generation.<br />

In cooperatively breeding, or ‘helping’, species,<br />

offspring from previous breeding seasons may stay with<br />

their parents and assist them in rearing additional siblings.<br />

Most helpers appear to help because no breeding opportunities<br />

are available elsewhere. A population census<br />

would mistakenly incorporate them in an analysis of population<br />

viability, when in fact the N e is much lower than it<br />

appears. To reduce the risk of extinction of the Seychelles<br />

warbler Acrocephalus sechellensis, which was restricted to a<br />

single island, Komdeur et al. (1995) translocated helpers to<br />

an unoccupied island. These helpers bred successfully and<br />

eventually accumulated helpers of their own when the new<br />

island’s territories became fully occupied. Thus knowledge<br />

of the adaptive function of cooperative breeding was used<br />

to found a new population of this endangered species without<br />

reducing the effective population size of the single<br />

source population. See also: Eusociality and Cooperation<br />

In other cases, wildlife management intended to protect<br />

endangered species has had unintended negative consequences<br />

on parental care. Poaching of rhinos for their<br />

horns resulted in drastic declines of African rhino species.<br />

In Namibia wildlife managers captured and removed the<br />

horns of their few remaining rhinos to make them worthless<br />

to horn-poachers. Unfortunately the calves of de-horned<br />

mother rhinos had low survival, probably due to the reduced<br />

ability of their mothers to protect their offspring<br />

from predators (Berger and Cunningham, 1994).<br />

ENCYCLOPEDIA OF LIFE SCIENCES & 2008, John Wiley & Sons, Ltd. www.els.net 5


Effects of Small Population Size on<br />

Behaviour<br />

The reduced reproductive success of individuals in small<br />

populations, called the ‘Allee Effect’, is of great concern to<br />

conservationists. Sometimes this reduced reproduction is<br />

simply a consequence of low population densities and the<br />

inability of individuals to find one another for mating. In<br />

other cases the adaptive benefits of living in social groups<br />

are reversed below a threshold group size. For example, in<br />

some normally social species it might be more costly to<br />

lifetime reproductive success to live in a too small group<br />

than to live an entirely solitary existence. Allee effects are<br />

not readily predicted if a ‘species typical’ viewpoint of behaviour<br />

is adopted, but will be anticipated by conservation<br />

behaviourists because they focus at the level of individual<br />

differences in decision-making behaviour. We explore several<br />

examples of the Allee effect in endangered mammal<br />

species.<br />

African wild dogs<br />

Small population sizes may have particularly profound<br />

effects on the behaviour and population viability of highly<br />

social animals such as the African wild dog, Lycaon pictus.<br />

Courchamp et al. (2002) demonstrated that small African<br />

wild dog populations may suffer reduced viability due to a<br />

trade-off between hunting and pup-guarding. Wild dog<br />

hunting is performed predominantly by the pack, and<br />

larger hunting groups are typically more successful at capturing<br />

prey. Additionally, although the majority of the<br />

pack is hunting, one adult usually remains to guard pups<br />

from predation. Thus, when packs reach a threshold level<br />

of fewer than about five individuals, the trade-off between<br />

hunting success and pup protection may have serious<br />

effects on population viability. Either hunting success or<br />

pup survival will be reduced at low group sizes. African<br />

wild dog populations composed of very small groups can be<br />

viable only if the risks of pup predation are also very low.<br />

Saiga antelope<br />

Behavioural shifts associated with reduced population size<br />

have also been demonstrated in the saiga antelope (Saiga<br />

tatarica) in Asia. Saiga populations have seen a catastrophic<br />

decrease in the number of males, due to an increase<br />

in poaching for saiga meat and horns. Poaching of adult<br />

males resulted in an extremely female-biased sex ratio,<br />

which in turn led to a decline in the number of pregnancies<br />

(Milner-Gulland, 2003). Normally, in the polygnous mating<br />

system used by saiga antelopes, males will defend a<br />

harem of 12<strong>–</strong>30 females. In female-biased populations,<br />

dominant females aggressively excluded subdominant females<br />

from mating with the few remaining males (Milner-<br />

Gulland, 2003). As a consequence first-year females in<br />

poached populations do not conceive, and the populations<br />

decline due to the direct mortality of males by poaching,<br />

6<br />

Conservation and Behaviour<br />

and the reduced fecundity of females due to a behavioural<br />

Allee effect.<br />

Hawaiian monk seals<br />

In contrast to the saiga, small populations of the endangered<br />

Hawaiian monk seal Monachus schauinslandi have<br />

unusually high numbers of adult males. Although the reasons<br />

for the male-biased sex ratios seen in this species are<br />

unknown, small population size may be a contributing<br />

factor (Starfield et al., 1995). Male-biased sex ratios are<br />

thought to increase the probability of mobbing, a form of<br />

intra-sexual competition in which individuals or groups of<br />

males may injure and/or kill adult females and immature<br />

pups during mating attempts. Mobbing exacerbates the<br />

male-biased sex ratios by further reducing the numbers of<br />

breeding females. Conservation behaviourists have effectively<br />

reduced mobbing by translocating surplus males to<br />

islands unoccupied by seals, allowing females to mate and<br />

rear pups more successfully.<br />

Managing Human<strong>–</strong>Wildlife Conflict<br />

Utilitarian value through sustainable use of wildlife by humans<br />

is thought to be an essential incentive for nature<br />

preservation (Hutton and Leader-Williams, 2003). An understanding<br />

of how and why animals behave the way they<br />

do will be instrumental in managing the population-level<br />

consequences of the direct and indirect effects of harvesting<br />

animals. Examples of wildlife (or their habitat) use by humans<br />

that benefit from a conservation behaviour approach<br />

include bycatch reduction in marine fisheries, management<br />

of ecotourism, reducing crop raiding by elephants and disruption<br />

of animal communication by pollution.<br />

Bycatch reduction<br />

An estimated 27 million tons of non-target species are discarded<br />

annually in fisheries worldwide (Alverson et al.,<br />

1994). Of greatest concern is the destruction of individuals<br />

of long-lived, slow breeding (also called ‘K-selected’) species<br />

such as porpoises, sea turtles and albatrosses, whose<br />

populations cannot recover quickly from high rates of<br />

adult mortality. Understanding how and why non-target<br />

species are being caught in these fisheries are intrinsically<br />

behavioural questions. The development of sensory deterrents,<br />

such as acoustic ‘pingers’ to warn porpoises away<br />

from gill nets (Kastelein et al., 2006), and olfactory repellants<br />

to keep sea birds from pursuing baited hooks in longline<br />

fisheries (Pierre and Norden, 2006) will require<br />

continued study to understand how their effectiveness is<br />

impacted by individual learning, and how evolutionary<br />

history determines which species respond to these bycatch<br />

reduction methods. Bycatch reduction devices that allow<br />

non-target species to exit trawl nets without reducing the<br />

capture of the target organism will require further investigation<br />

of the sensory world of these species and the<br />

ENCYCLOPEDIA OF LIFE SCIENCES & 2008, John Wiley & Sons, Ltd. www.els.net


decision rules they use in response to those stimuli. Differential<br />

light preferences of target and non-target species may<br />

also be useful in developing more selective fisheries<br />

(Marchesan et al., 2005). See also: Reproductive Strategies<br />

Ecotourism<br />

Generally, ecotourism is thought to change wildlife behaviour,<br />

but some studies have been able to link human disturbance<br />

with positive effects on animal populations. for<br />

example, the presence of bear viewing tours in British Columbia<br />

temporarily displaces adult male brown bears<br />

U. arctos, which in turn allows for females with cubs to<br />

spend more time fishing and foraging due to the fact that<br />

female bears appear to be less vigilant around humans than<br />

adult males. The displacement of large males by the presence<br />

of ecotourists is thought to create a temporal refuge<br />

that enhances feeding opportunities for subordinate age/<br />

sex classes (Nevin and Gilbert, 2005). More commonly investigators<br />

conclude that human disturbance has a neutral<br />

or only slightly negative effect on wildlife. Minimal reaction<br />

to human observers occurs in eastern chimpanzees<br />

(Pan troglodytes schweinfurthi) in the Kibale Forest,<br />

Uganda, provided that tourists visit in small groups<br />

(Johns, 1996). Magellanic penguins Spheniscus magellanicus<br />

in Argentina show both behavioural and physiological<br />

habituation to tourist visitation (Walker et al.,<br />

2006). It is not known, however, if the ontogenetic habituation<br />

of these penguins serves an adaptive function. What<br />

if becoming tamer around humans puts penguins at greater<br />

risk of ignoring other predators? Inter-specific comparisons<br />

of stress responses to disturbance of penguin species<br />

that evolved in the absence of native terrestrial predators,<br />

such as the Galapagos penguin Spheniscus mendiculus (Figure<br />

2), with those that show fearful reactions would be<br />

helpful to predict the conservation consequences of expanding<br />

ecotourism to additional penguin species. The<br />

challenge of ecotourism has always been to combine the<br />

demands of tourists with the needs of local populations and<br />

the conservation of protected areas. Not all behavioural<br />

change by animals visited by ecotourists is harmful to<br />

population viability. Sustainable ecotourism requires that<br />

conservationists minimize disturbance that impairs the<br />

lifetime reproductive success of individual animals.<br />

Crop raiding by elephants<br />

Human<strong>–</strong>elephant conflict over crop raiding is a significant<br />

threat to elephant conservation efforts in Africa and Asia.<br />

In addition to imposing significant economic burdens on<br />

humans that farm near elephant reserves, elephants may<br />

harm or even kill humans who attempt to repel them.<br />

Conservation behaviourists are attempting to properly understand<br />

the proximate and ultimate factors that promote<br />

crop raiding. The benefits of crop raiding to the elephants<br />

are dependent on the proximity of natural habitats to the<br />

crop fields, the timing of crop ripening and the spatial distribution<br />

of wild forage items (Rode et al., 2006). Males<br />

between the ages of 10 and 14 years and between the ages of<br />

20 and 24 years are associated with crop raiding, potentially<br />

due to three factors: social independence, risk tolerance<br />

and male<strong>–</strong>male competition (Chiyo et al., 2005). In<br />

addition to practical methods for preventing elephants<br />

from reaching crop fields (such as trenches, electric fences<br />

and grass buffers), behaviourists must develop management<br />

techniques that co-opt male elephant strategies that<br />

promote conflict with their human neighbours.<br />

Pollution<br />

Conservation and Behaviour<br />

Figure 2 Sustainable ecotourism provides financial incentives to protect<br />

natural areas. Conservation behaviour studies are needed to determine if<br />

island species, such as this Galapagos penguin, habituate more easily to<br />

human disturbance because they have evolved without terrestrial predators.<br />

Photo courtesy of Clifford Ochs.<br />

Human activities contaminate nearby animal habitats with<br />

a wide range of pollutants that may interfere with the<br />

communication systems of animal species. In the presence<br />

of such information disruptors, organisms may fail to detect<br />

important environmental cues required for the acquisition<br />

of food, mates, habitats and predators. For example,<br />

pesticides and heavy metals may interfere with antipredator<br />

responses (Lu¨ rling and Scheffer, 2007). Excess<br />

artificial light (light pollution) disrupts orientation cues in<br />

hatchling marine turtles (Witherington, 1997) and other<br />

nocturnally active wildlife. The nocturnal beach mouse is<br />

unable to forage efficiently in areas illuminated by artificial<br />

lights (Bird et al., 2004). Man-made noises, also serve as<br />

ENCYCLOPEDIA OF LIFE SCIENCES & 2008, John Wiley & Sons, Ltd. www.els.net 7


stressors to animal populations causing increased hypothalamic<br />

response resulting in physiological stress that may<br />

cause dispersal or impair reproduction.<br />

Implications for Management<br />

Problems of animal behaviour occur in most areas of conservation,<br />

and conservation behaviour is likely to become<br />

an even more pertinent tool to prevent biodiversity loss as<br />

wildlife becomes restricted to parks and preserves surrounded<br />

by anthropogenic environments (Buchholz,<br />

2007). Unfortunately most conservation biologists and<br />

wildlife managers have little training in the ethological<br />

framework or methodology that underpins conservation<br />

behaviour. Despite calls for the inclusion of behaviourists<br />

on conservation planning teams (Arcese et al., 1997), conservation<br />

behaviour remains poorly integrated with conservation<br />

biology (Caro, 2007; Angeloni et al., 2008). Our<br />

ability to protect biodiversity despite global climate<br />

change, exponential human population growth and unsustainable<br />

resource use will require that ethologists participate<br />

in conservation management, and demonstrate more<br />

effectively that the mechanisms, ontogeny, adaptive function<br />

and evolutionary history of animal behaviour have<br />

practical import to conserving nature.<br />

References<br />

Ahlering MA and Faaborg J (2006) Avian habitat management<br />

meets conspecific attraction: if you build it, will they come? Auk<br />

23: 301<strong>–</strong>312.<br />

Alberts AC, Lemm JM, Perry AM, Morici LA and Phillips JA<br />

(2002) Temporary alteration of local social structure in a<br />

threatened population of Cuban iguanas (Cyclura nubila). Behavioral<br />

Ecology and Sociobiology 51: 324<strong>–</strong>335.<br />

Alverson DL, Freeberg MH, Murawski SA and Pope JG (1994)<br />

A global assessment of fisheries bycatch and discards. FAO<br />

Fisheries Technical Paper 339, Food and Agriculture Organization<br />

of the United Nations, Rome, http://www.fao.org/docrep/003/T4890E/T4890E00.HTM<br />

Angeloni L, Schlaepfer MA, Lawler JJ and Crooks KR (2008)<br />

A reassessment of the interface between conservation and behaviour.<br />

Animal Behaviour 75: 731<strong>–</strong>737.<br />

Arcese P, Keller LF and Cary JR (1997) Why hire a behaviorist into<br />

a conservation or management team? In: Clemmons JR and<br />

Buchholz R (eds) Behavioral Approaches to Conservation in the<br />

Wild, pp. 48<strong>–</strong>71. Cambridge, UK: Cambridge University Press.<br />

Battini J (2004) When good animals love bad habitats: ecological<br />

traps and the conservation of animal populations. Conservation<br />

Biology 18: 1482<strong>–</strong>1491.<br />

Beissinger SR (1997) Integrating behavior into conservation<br />

biology: potentials and limitations. In: Clemmons JR and<br />

Buchholz R (eds) Behavioral Approaches to Conservation in the<br />

Wild, pp. 23<strong>–</strong>47. Cambridge, UK: Cambridge University Press.<br />

Berger J and Cunningham C (1994) Active intervention and conservation:<br />

Africa’s pachyderm problem. Science 263:<br />

1241<strong>–</strong>1242.<br />

8<br />

Conservation and Behaviour<br />

Bird B, Branch L and Miller D (2004) Effects of coastal lighting<br />

on foraging behavior of beach mice. Conservation Biology<br />

18: 1435<strong>–</strong>1439.<br />

Bolger DT, Newmark WD, Morrison TA and Doak DF (2008)<br />

The need for integrative approaches to understand and conserve<br />

migratory ungulates. Ecology Letters 11: 63<strong>–</strong>77.<br />

Buchholz R (2007) Behavioural biology: an effective and<br />

relevant conservation tool. Trends in Ecology and Evolution<br />

22: 401<strong>–</strong>407.<br />

Canadian Wildlife Service and U.S. Fish and Wildlife Service<br />

(2005) <strong>International</strong> Recovery Plan for the Whooping Crane,<br />

162pp. Ottawa: Recovery of Nationally Endangered Wildlife<br />

(RENEW), and Albuquerque, NM: U.S. Fish and Wildlife<br />

Service.<br />

Caro T (2007) Behavior and conservation: a bridge too far? Trends<br />

in Ecology and Evolution 22: 394<strong>–</strong>400.<br />

Castello´ n TD and Sieving KE (2005) An experimental test of matrix<br />

permeability and corridor use by an endemic understory<br />

bird. Conservation Biology 20: 135<strong>–</strong>145.<br />

Chiyo P and Cochrane E (2005) Population structure and behavior<br />

of crop-raiding elephants in Kibale National Park,<br />

Uganda. African Journal of Ecology 43: 233<strong>–</strong>241.<br />

Clevenger AP and Waltho N (2005) Performance indices to identify<br />

attributes of highway crossing structures facilitating<br />

movement of large mammals. Biological Conservation 121:<br />

453<strong>–</strong>464.<br />

Clout MN, Elliott GP and Robertson BC (2002) Effects of supplementary<br />

feeding on the offspring sex ratio of kakapo: a dilemma<br />

for the conservation of a polygynous parrot. Biological<br />

Conservation 107: 13<strong>–</strong>18.<br />

Coltman DW, O’Donoghue P, Jorgenson JT et al. (2003) Undesirable<br />

evolutionary consequences of trophy hunting. Nature<br />

426: 655<strong>–</strong>658.<br />

Courchamp F, Rasmussen GSA and Macdonald DW (2002)<br />

Small pack size imposes a trade-off between hunting and pupguarding<br />

in the painted hunting dog Lycaon pictus. Behavioral<br />

Ecology 13: 20<strong>–</strong>27.<br />

Cryan PM and Brown AC (2007) Migration of bats past a remote<br />

island offers clues toward the problem of bat fatalities at wind<br />

turbines. Biological Conservation 139: 1<strong>–</strong>11.<br />

Davis LA, Roalson EH, Cornell KL, McClanahan KD and<br />

Webster MS (2006) Genetic divergence and migration patterns<br />

in a North American passerine bird: implications for evolution<br />

and conservation. Molecular Ecology 15: 2141<strong>–</strong>2152.<br />

Diego-Rasilla FJ and Luengo RM (2004) Heterospecific call recognition<br />

and phonotaxis in the orientation behavior of the<br />

marbled newt, Triturus marmoratus. Behavioral Ecology and<br />

Sociobiology 55: 556<strong>–</strong>560.<br />

Doherty PF Jr, Sorci G, Royle JA et al. (2003) Sexual selection<br />

affects local extinction and turnover in bird communities. Proceedings<br />

of the National Academy of Sciences of the USA 100:<br />

5858<strong>–</strong>5862.<br />

Fisher HS, Swaisgood RR and Fitch-Snyder H (2003) Odor familiarity<br />

and female preferences for males in a threatened primate,<br />

the pygmy loris Nycticebus pygmaeus: applications for<br />

genetic management of small populations. Naturwissenschaften<br />

90: 509<strong>–</strong>512.<br />

Fletcher RJ Jr (2007) Species interactions and population density<br />

mediate the use of social cues for habitat selection. Journal of<br />

Animal Ecology 76: 598<strong>–</strong>606.<br />

ENCYCLOPEDIA OF LIFE SCIENCES & 2008, John Wiley & Sons, Ltd. www.els.net


Gagnon JW, Theimer TC, Dodd NL, Boe S and Schweinsburg RE<br />

(2007) Traffic volume alters elk distribution and highway crossings<br />

in Arizona. Journal of Wildlife Management 71: 2318<strong>–</strong>2323.<br />

Gilroy JJ and Sutherland WJ (2007) Beyond ecological traps:<br />

perceptual errors and undervalued resources. Trends in Ecology<br />

and Evolution 22: 351<strong>–</strong>356.<br />

Haddad NM (1999) Corridor use predicted from behaviors at<br />

habitat boundaries. American Naturalist 153: 215<strong>–</strong>227.<br />

Hutton JM and Leader-Williams N (2003) Sustainable use and<br />

incentive-driven conservation: realigning human and conservation<br />

interests. Oryx 37: 215<strong>–</strong>226.<br />

Johns BG (1996) Responses of chimpanzees to habituation and<br />

tourism in the Kibale Forest, Uganda. Biological Conservation<br />

78: 257<strong>–</strong>262.<br />

Kastelein RA, Jennings N, Verboom WC, de Haan D and Schooneman<br />

NM (2006) Differences in the response of a striped<br />

dolphin (Stenella coeruleoalba) and a harbour porpoise<br />

(Phocoena phocoena) to an acoustic alarm. Marine Environmental<br />

Research 61: 363<strong>–</strong>378.<br />

Komdeur J, Castle G, Huffstad A et al. (1995) Transfer experiments<br />

of Seychelles warblers to new islands: changes in dispersal<br />

and helping behaviour. Animal Behaviour 49: 695<strong>–</strong>708.<br />

Laurenson MK, Wielebnowski N and Caro TM (1995) Extrinsic<br />

factors and juvenile mortality in cheetahs. Conservation Biology<br />

9: 1329<strong>–</strong>1331.<br />

Lu¨ rling M and Scheffer M (2007) Info-disruption: pollution and<br />

the transfer of chemical information between organisms. Trends<br />

in Ecology and Evolution 22(7): 374<strong>–</strong>379.<br />

Marchesan M, Spoto M, Verginella L and Ferrero EA (2005)<br />

Behavioural effects of artificial light on fish species of commercial<br />

interest. Fisheries Research 73: <strong>17</strong>1<strong>–</strong>185.<br />

Mata C, Hervás I, Herranz J, Suárez F and Malo JE (2005) Complementary<br />

use by vertebrates of crossing structures along a<br />

fenced Spanish motorway. Biological Conservation 124: 397<strong>–</strong>405.<br />

McDonald RA, Delahay RJ, Carter SP, Smith GC and Cheeseman<br />

CL (2008) Perturbing implications of wildlife ecology for<br />

disease control. Trends in Ecology and Evolution 23: 53<strong>–</strong>56.<br />

Milner-Gulland EJ (2003) Reproductive collapse in saiga antelope<br />

harems. Nature 422: 135.<br />

Nevin OT and Gilbert BK (2005) Perceived risk, displacement and<br />

refuging in brown bears: positive impacts of ecotourism? Biological<br />

Conservation 121: 611<strong>–</strong>622.<br />

Ng SJ, Dole JW, Sauvajot RM, Riley SPD and Valone TJ (2004)<br />

Use of highway undercrossings by wildlife in southern California.<br />

Biological Conservation 115: 499<strong>–</strong>507.<br />

Parker PG and Waite TA (1997) Mating systems, effective<br />

population size, and conservation of natural populations. In:<br />

Clemmons JR and Buchholz R (eds) Behavioral Approaches to<br />

Conservation in the Wild, pp. 243<strong>–</strong>261. Cambridge, UK: Cambridge<br />

University Press.<br />

Pierre JP and Norden WS (2006) Reducing seabird bycatch in<br />

longline fisheries using a natural olfactory deterrent. Biological<br />

Conservation 130: 406<strong>–</strong>415.<br />

Quader S (2005) Mate choice and its implications for conservation<br />

and management. Current Science 89: 1220<strong>–</strong>1229.<br />

Rittenhouse TAG and Semlitsch RD (2006) Grasslands as movement<br />

barriers for a forest-associated salamander: migration<br />

behavior of adult and juvenile salamanders at a distinct habitat<br />

edge. Biological Conservation 131: 14<strong>–</strong>22.<br />

Rode K, Chiyo P, Chapman C and McDowell L (2006) Nutritional<br />

ecology of elephants in Kibale National Park, Uganda,<br />

and its relationship with crop-raiding behavior. Journal of<br />

Tropical Ecology 22: 441<strong>–</strong>449.<br />

Rosenberg DK, Noon BR, Megahan JW and Meslow EC (1998)<br />

Compensatory behavior of Ensatina eschscholtzii in biological<br />

corridors: a field experiment. Canadian Journal of Zoology<br />

76: 1<strong>17</strong><strong>–</strong>133.<br />

Schlaepfer MA, Runge MC and Sherman PW (2002) Ecological<br />

and evolutionary traps. Trends in Ecology and Evolution<br />

<strong>17</strong>: 474<strong>–</strong>480.<br />

Starfield AM, Roth JD and Ralls K (1995) Mobbing in Hawaiian<br />

monk seals (Monachus schauinlandi): the value of simulation<br />

modeling in the absence of apparently crucial data. Conservation<br />

Biology 9: 166<strong>–</strong><strong>17</strong>4.<br />

Thomas GH, Lanctot RB and Sze´ kely T (2006) Can intrinsic factors<br />

explain population decline in North American breeding<br />

shorebirds? A comparative analysis. Animal Conservation<br />

9: 252<strong>–</strong>258.<br />

Walker BG, Boersma PD and Wingfield JC (2006) Habituation of<br />

adult magellanic penguins to human visitation as expressed<br />

through behavior and corticosterone secretion. Conservation<br />

Biology 20: 146<strong>–</strong>154.<br />

Ward MP and Schlossberg S (2004) Conspecific attraction and the<br />

conservation of territorial songbirds. Conservation Biology<br />

18: 519<strong>–</strong>525.<br />

Whiting SD, Murray W, Macrae I et al. (2008) Non-migratory<br />

breeding by isolated green sea turtles (Chelonia mydas) in the<br />

Indian Ocean: biological and conservation implications.<br />

Naturwissenschaften 95: 355<strong>–</strong>360.<br />

Whitman K, Starfield AM, Quadling HS and Packer C (2004)<br />

Sustainable trophy hunting of African lions. Nature 428:<br />

<strong>17</strong>5<strong>–</strong><strong>17</strong>8.<br />

Witherington BE (1997) The problem of photopollution for sea<br />

turtles and other nocturnal animals. In: Clemmons JR and<br />

Buchholz R (eds) Behavioral Approaches to Conservation in the<br />

Wild, pp. 303<strong>–</strong>328. Cambridge, UK: Cambridge University<br />

Press.<br />

Zedrosser A, Støen O, Saebø S and Swenson JE (2007) Should I<br />

stay or should I go? Natal dispersal in the brown bear. Animal<br />

Behaviour 74: 369<strong>–</strong>376.<br />

Further Reading<br />

Conservation and Behaviour<br />

Caro T (1998) Behavioral Ecology and Conservation Biology. New<br />

York: Oxford University Press, New York.<br />

Festa-Bianchet M and Apollonio M (2003) Animal Behavior and<br />

Wildlife Conservation. Washington, DC: Island Press.<br />

Gosling LM and Sutherland WJ (2000) Behaviour and Conservation.<br />

New York: Cambridge University Press.<br />

Harcourt AH and Stewart KJ (2007) Socio-ecology and gorilla<br />

conservation. Chap. 14. In: Gorilla Society. Chicago: University<br />

of Chicago Press.<br />

Sutherland WJ (1996) From Individual Behaviour to Population<br />

Ecology. Oxford Series in Ecology and Evolution. New York:<br />

Oxford University Press.<br />

ENCYCLOPEDIA OF LIFE SCIENCES & 2008, John Wiley & Sons, Ltd. www.els.net 9


Behavioural biology: an effective and<br />

relevant conservation tool<br />

Richard Buchholz<br />

Opinion TRENDS in Ecology and Evolution Vol.22 No.8<br />

Department of Biology, University of Mississippi, University, MS 38677-1848, USA<br />

‘Conservation behaviour’ is a young discipline that<br />

investigates how proximate and ultimate aspects of<br />

the behaviour of an animal can be of value in preventing<br />

the loss of biodiversity. Rumours of its demise are<br />

unfounded. Conservation behaviour is quickly building<br />

a capacity to positively influence environmental decision<br />

making. The theoretical framework used by animal behaviourists<br />

is uniquely valuable to elucidating integrative<br />

solutions to human-wildlife conflicts, efforts to reintroduce<br />

endangered species and reducing the deleterious<br />

effects of ecotourism. Conservation behaviourists must<br />

join with other scientists under the multidisciplinary<br />

umbrella of conservation biology without giving up on<br />

their focus: the mechanisms, development, function and<br />

evolutionary history of individual differences in behaviour.<br />

Conservation behaviour is an increasingly relevant<br />

tool in the preservation of nature.<br />

The origins of conservation behaviour<br />

Behavioural biology did not rank among the fields included<br />

under the multidisciplinary umbrella of conservation<br />

biology when this new science was created in 1985 [1].<br />

Perhaps as a consequence, behavioural study was not<br />

incorporated into the first conservation biology textbooks<br />

and conservation was not mentioned in the animal behaviour<br />

texts of the era [2]. It required another decade for the<br />

nascent field of ‘conservation behaviour’ to coalesce from<br />

symposia, workshops and the resulting multi-authored<br />

volumes that appeared after the mid-1990s (citations in<br />

[3]). Some have suggested that behavioural biologists<br />

became interested in conservation only because they<br />

anticipated a new source of research funding [3,4], but<br />

those of us who were graduate students at the time know<br />

the real reason that conservation behaviour arose: in the<br />

face of biodiversity loss and widespread habitat destruction,<br />

we wanted our science to be relevant to saving the<br />

natural world.<br />

Wishing for behavioural biology to be useful to<br />

conservation, however, is not evidence that it is. Indeed,<br />

some early proponents of conservation behaviour recently<br />

seem to have lost their faith in the discipline * . Likewise, I<br />

have observed that some aspiring conservation behaviourists<br />

are questioning the success of conservation behaviour<br />

at integrating with mainstream conservation efforts. Their<br />

Corresponding author: Buchholz, R. (byrb@olemiss.edu).<br />

* Caro, T. (2006) Challenges and opportunities for behavioural ecologists to save<br />

Planet Earth. 25 July Plenary session, 11th Congress of the <strong>International</strong> Society for<br />

Behavioral Ecology, Tours, France.<br />

Available online 27 June 2007.<br />

www.sciencedirect.com 0169-5347/$ <strong>–</strong> see front matter ß 2007 Elsevier Ltd. All rights reserved. doi:10.1016/j.tree.2007.06.002<br />

pessimism is in sharp contrast to my own view: a decade<br />

after I co-edited the first book on the subject [5], Iam<br />

pleased with recent developments in the adolescence of<br />

conservation behaviour. After only a decade of existence, it<br />

is premature to dismiss the relevancy of the conservation<br />

behaviourist to saving biodiversity. Here, I show evidence<br />

of the vibrancy of this growing field (Box 1) and then<br />

recount how the theoretical framework of conservation<br />

behaviourists is already positioning them to help solve<br />

the types of conservation issues that will be especially<br />

vexing in the coming decades.<br />

Defining conservation behaviour<br />

It is obvious that species-typical patterns of animal<br />

movement, feeding and mating must be considered in<br />

conservation planning. Thus, the debate over the role of<br />

modern animal behaviour studies in conservation is not a<br />

question of whether major differences in species comportment<br />

(e.g. seasonal migration versus year-round residency)<br />

are important to conservation planning. Instead,<br />

it is a disagreement over whether the discipline-specific<br />

training of animal behaviourists, sensu stricto, is a valuable<br />

addition to conservation action teams [6]. I have<br />

observed two possible reasons why conservation behaviour<br />

has not made major in-roads in traditional conservation<br />

biology circles: conservation ecologists believe that they<br />

already ‘do’ behaviour, and many think that animal behaviourists<br />

work at scales of minor value to protecting entire<br />

landscapes, the most cost-effective means of conservation.<br />

Behaviour thinking<br />

On the surface, ‘doing’ behaviour seems relatively easy. For<br />

example, conservation biologists might walk through a<br />

forest quantifying territorial vocalizations to census a<br />

population of songbirds. Such superficial uses of behavioural<br />

biology are no doubt useful, but they, like<br />

species-typical behavioural descriptions by natural historians<br />

[7], are not conservation behaviour. Conservation<br />

behaviour takes advantage of an investigatory framework<br />

implicit to all modern animal behaviour studies. This<br />

framework is commonly referred to as Tinbergen’s ‘four<br />

questions’, because it was first proposed by Nobel-Prizewinning<br />

ethologist Niko Tinbergen as a way to guide<br />

behavioural research [8]. Tinbergen suggested that we<br />

can ask four mutually exclusive questions about any one<br />

behaviour by considering both proximate and ultimate<br />

explanations for the cause and origin of that behaviour<br />

pattern (see Table I in Box 2). Some refer to this use of<br />

Tinbergen’s framework as ‘behaviour thinking’, and to


402 Opinion TRENDS in Ecology and Evolution Vol.22 No.8<br />

Box 1. The influence of conservation behaviour is growing<br />

This tenth anniversary of the publication of the first books on<br />

conservation behaviour [5,44] is an arbitrary date on which to assess<br />

this young discipline’s success at integrating with conservation<br />

biology. Rather than expecting conservation behaviour to be fully<br />

fledged at this point, we should look for signs that it continues to<br />

strengthen in terms of training resources, pervasiveness in the<br />

published literature, and acceptance by grant-awarding agencies<br />

and career mentors. Supportive evidence includes the following.<br />

Eighty percent of recently published animal behaviour texts<br />

address conservation in some fashion [2]. Page content that<br />

contains reference to conservation now averages 2% for texts<br />

published during the past five years, compared with 0% in the<br />

decades before that.<br />

Conservation behaviour training resources are readily available<br />

on the Internet (http://www.animalbehavior.org/Committees/<br />

Conservation). It is now easier for aspiring conservation behaviourists<br />

(and their mentors) to access background information,<br />

including a general bibliography, funding sources and graduate<br />

programs. Also available for download is ‘The Conservation<br />

Behaviorist’, a journal written for both behaviourists and nonscientist<br />

decision makers.<br />

The use of behavioural biology in published conservation studies<br />

is not uncommon; in fact, it is growing. Linklater [45] discovered<br />

that the percentage of conservation publications that mention<br />

behaviour has increased nearly threefold since the conception of<br />

conservation behaviour in the early 1990s.<br />

The EO Wilson Student Conservation Research Grant Award of<br />

the Animal Behavior Society (ABS) was created to provide an<br />

annual small grant to foster the career development of conservation<br />

behaviourists. Conservation grant awardees have<br />

studied logging effects on tree shrew mating systems, stream<br />

pollution interference with fish communication and dragonfly<br />

oviposition requirements as bioindicators of wetland quality.<br />

Perhaps more significant than these new conservation behaviour<br />

research funds are the types of research that ‘regular’ student<br />

behaviourists are doing. In 2006, 15% of the standard Student<br />

Research Awards from the ABS were for projects with conservation<br />

themes.<br />

The ‘father’ of conservation biology, Michael Soulé, said that<br />

the new discipline: ‘‘should attract and penetrate every field<br />

that could possibly benefit and protect the diversity of life’’ [46].<br />

These demographics suggest that the population of conservation<br />

behaviourists is indeed growing.<br />

proximate and ultimate explanations as ‘how’ and ‘why’<br />

questions.<br />

Proximate questions about behaviour consider how<br />

an individual is able to perform an activity. They ask<br />

about the mechanisms within an organism that make it<br />

possible for it to behave in a certain way. The proximate<br />

causes of behaviour include the sensory and endocrine<br />

mechanisms that regulate behaviour. However, we know<br />

that these mechanisms can be modified by individual<br />

experience; thus, we must consider the proximate origins<br />

of behaviour as well (i.e. how learning modifies behaviour).<br />

Ultimate questions about behaviour, on the other hand,<br />

ask why animal species have evolved the proximate systems<br />

that enable them to behave the way they do. The<br />

ultimate cause of a behaviour must explain how it helps the<br />

individual survive and reproduce. If we ponder the ultimate<br />

origin of that same behaviour pattern, we are examining<br />

its evolutionary history by comparing how that<br />

behaviour differs across a group of related species. Tinbergen’s<br />

framework is most easily explained by applying it to a<br />

behavioural example (Box 2).<br />

www.sciencedirect.com<br />

Box 2. Understanding Tinbergen’s framework for studying<br />

behaviour<br />

Tinbergen’s four complementary approaches to studying a behaviour<br />

pattern are implicit to modern animal behaviour studies (see<br />

Table I). This framework is most easily explained by applying it to an<br />

example of human behaviour: automobile drivers stop their cars<br />

when a traffic light turns red.<br />

Mechanisms<br />

We can ask how automobile drivers are able to perceive the red<br />

colour of the traffic light, the pattern of neural depolarization that<br />

allows a decision to be made in the central nervous system and how<br />

that decision is conveyed to the muscles that control the placement<br />

of the foot on the brake. The problem of ‘road rage’ suggests that<br />

latency to apply the brake may be affected by the hormonal milieu of<br />

the driver.<br />

Ontogeny<br />

Humans are not born knowing how to drive an automobile. We can<br />

ask questions about how humans go about learning to stop at a red<br />

light and how it is that development affects the ability to learn. For<br />

example, accident statistics suggest that teenagers and senior<br />

citizens may have greater difficulty stopping at red lights than other<br />

drivers.<br />

Function<br />

We can investigate why stopping at red lights allows drivers to live<br />

longer and reproduce successfully. The legal and financial costs and<br />

benefits that contribute to the individual fitness of drivers are<br />

worthy of study. We might also consider why some individuals are<br />

more likely to cooperate with other drivers in their society.<br />

Phylogeny<br />

An investigation of the history of traffic signals might tell us how<br />

responses to red lights have changed over time. In humans, we can<br />

look at how populations have independently developed means of<br />

making intersections safer. If we compare humans to other species,<br />

we might ask how the colour red has evolved as a warning signal.<br />

Table I. Tinbergen’s framework for organizing the study of<br />

behaviour proposes that any behaviour pattern should be<br />

investigated from four complementary perspectives<br />

Cause Origin<br />

Proximate Mechanisms Ontogeny<br />

Ultimate Function Phylogeny<br />

The beauty of Tinbergen’s four questions is that they<br />

force us to consider multiple, complementary explanations<br />

for the same behaviour [9]. In terms of saving biodiversity,<br />

this framework is especially effective in situations in which<br />

the behavioural adaptations of wildlife are at odds with<br />

anthropogenic landscapes [10]. Conservationists might<br />

eventually stumble upon the need for knowing both<br />

the ‘how’ and the ‘why’ of animal behaviour (Box 3), but<br />

it would be much more cost effective and time efficient<br />

if Tinbergen’s framework was applied at the onset of<br />

conservation research.<br />

Where does conservation behaviour fit?<br />

It would be preposterous to claim that behavioural biology<br />

is the conservation ‘cure all’. Behavioural study is neither<br />

appropriate nor a priority at all levels of conservation<br />

action (Table 1 and [11]). The need for behaviourists will<br />

be greatest when we are confronted with the challenge of<br />

maintaining animals marooned in protected islands of<br />

habitat, isolated in a sea of humanity, and managing


Box 3. Harbour porpoises and gill nets: application of<br />

Tinbergen’s framework<br />

In 1994 y , Marquez and I used the example of a ‘real-world’<br />

conservation problem, the drowning of harbour porpoises Phocoena<br />

phocoena following entanglement in the gill nets of commercial<br />

fisherman, to explain how Tinbergen’s four questions could guide<br />

conservation research. Read and colleagues [47] had called for<br />

improved understanding of the behaviour of harbour porpoises<br />

around gill nets. Incidental bycatch of harbour porpoises in the US<br />

Gulf of Maine groundfish gill net fishery alone averaged over 2000<br />

individuals per year in the early 1990s, more than twice the<br />

allowable take rate.<br />

For heuristic purposes, we suggested that behavioural aspects<br />

of gill net bycatch research could be approached simultaneously<br />

from mechanistic, ontogenetic, functional and phylogenetic<br />

perspectives. Revisiting the problem now, it is rewarding to see<br />

that anti-entanglement research evolved within Tinbergen’s<br />

framework.<br />

Harbour porpoises appear to be in the vicinity of nets because<br />

nets are placed where porpoise prey are abundant [48]. Making nets<br />

from material that is more reflective of porpoise sonar did reduce<br />

bycatch significantly, but perhaps only because the nets were stiffer<br />

and less likely to entangle the porpoises [49]. Adding acoustic<br />

alarms, called ‘pingers’, to gill nets reduces the bycatch mortality of<br />

harbour porpoises [50]. Harbour porpoises show aversive reactions<br />

to pinging, including spatial avoidance and increasing respiration<br />

rate [51].<br />

Other small cetaceans, however, do not necessarily react the<br />

same way [51]. In a field experiment, bottlenose dolphins Tursiops<br />

truncatus (Figure I) appeared to use fish caught in the gill net as a<br />

food source, but dolphin groups were less likely to approach the<br />

net’s ‘zone of vulnerability’ when the pingers were activated [50].<br />

There is considerable interest in investigating how these species will<br />

react to long-term use of acoustic alarms. Will cetaceans become<br />

sensitized to pingers and avoid them more often, or will pingers<br />

become a ‘dinner bell’ of sorts, attracting porpoises and dolphins to<br />

pre-caught fish?<br />

Although the gill net responses of relatively few cetaceans<br />

have been studied so far, some researchers have hypothesized<br />

that vulnerability to natural predators, such as sharks and killer<br />

whales Orcinus orca, might explain species differences in the<br />

efficacy of using acoustic alarms to reduce incidental bycatch<br />

mortality.<br />

I do not mean to suggest that these studies are the result of our<br />

Society for Conservation Biology poster; these works developed<br />

organically from the need to solve the gill net entanglement<br />

problem. Nevertheless, this conservation problem demonstrates<br />

how a priori use of a conservation behaviour framework would<br />

foster an efficient conservation research plan [52].<br />

Figure I. Bottlenose dolphins are attracted to the fish caught in commercial gill<br />

nets (photo by Jill Frank).<br />

y Marquez, M. and Buchholz, R. (1994) An Ethological Framework for<br />

Conservation Biology. Poster at the annual meeting of the Society for<br />

Conservation Biology, University of Guadalajara, Jalisco, Mexico.<br />

www.sciencedirect.com<br />

Opinion TRENDS in Ecology and Evolution Vol.22 No.8 403<br />

the inevitable conflicts between the daily needs of humans<br />

and other animals. It is not pleasant to think of a future in<br />

which most of nature is drastically altered by humans, but<br />

the reality is undeniable [12].<br />

Global warming<br />

Conservation behaviourists are developing predictive tools<br />

for understanding which species in pristine communities<br />

will need behavioural management when their habitats<br />

are altered by direct and indirect human disturbance. Paz<br />

y Miño [13] has termed these circumstances ‘behavioural<br />

unknowns’ (after Myers ‘environmental unknowns’ [14]).<br />

The most pressing of these, perhaps, is the need to understand<br />

how the mechanisms, ontogeny, adaptiveness and<br />

phylogenetic diversity of animal behaviour will respond to<br />

climate change due to global warming. Conservation behaviourists<br />

have already begun to develop a body of literature<br />

that addresses behavioural responses to rapid climatic<br />

alterations (Box 4).<br />

Restoring balance to ecosystems<br />

Other behavioural unknowns may have less to do with how<br />

we are destroying habitat and more to do with our attempts<br />

to restore ecological integrity to human-altered landscapes.<br />

The reintroduction of large carnivores appears to<br />

enhance ecosystem biodiversity and stability [15]. Nonetheless,<br />

game managers fear that huntable (by humans)<br />

prey species will be decimated because of their naiveté<br />

after generations without non-human predation. By experimentally<br />

investigating the anti-predator responses of<br />

ungulates to predator cues before and after carnivore<br />

reintroduction, Berger [16] found that prey typically return<br />

to ‘normal’ anti-predator strategies within one generation<br />

of carnivore return. If politically feasible, conservation<br />

behaviour data such as these will be used to support the<br />

repatriation of carnivores so that balanced ecosystems are<br />

restored.<br />

Ecological prediction is a mainstay of traditional<br />

conservationists [<strong>17</strong>]. Game theory, and other skills in<br />

the behaviourist’s toolbox that take advantage of individual<br />

differences in behaviour, are currently being used to<br />

anticipate the population consequences of management<br />

options [18]. In addition to predicting how animals will<br />

respond to anthropogenic disturbances, animal behaviourists<br />

are influencing conservation management directly.<br />

Conservation behaviourists in action<br />

To date, the contributions of conservation behaviourists<br />

are much more than theoretical. Conservation behaviourists<br />

are already involved in hands-on efforts to restore and<br />

protect animal species. Here, I briefly review some recent<br />

contributions of note.<br />

Species reintroduction<br />

Although the future of biodiversity is in the wild, captive<br />

breeding of endangered species is sometimes an irreplaceable<br />

component of the conservationist’s toolbox [11,19].<br />

Conservation behaviourists have concentrated on two<br />

important aspects of captive propagation of endangered<br />

species: preventing ‘captive selection’ (i.e. maladaptive<br />

heritable changes in behaviour), and behavioural training


404 Opinion TRENDS in Ecology and Evolution Vol.22 No.8<br />

Table 1. Behavioural biology is relevant to multiple conservation contexts<br />

Conservation context Conservation tool Example of use<br />

Preventing biodiversity loss Reserve design a<br />

[37]<br />

Ecosystem management b<br />

[15]<br />

Population viability analysis b<br />

[43]<br />

Compromises with economic development Sustainable use b<br />

[29]<br />

Species and habitat restoration Field recovery of endangered species a<br />

[23]<br />

Captive breeding and reintroduction a<br />

[20]<br />

Ecosystem restoration a<br />

a<br />

Tools with current behaviour usage.<br />

[16]<br />

b<br />

Tools that will need more behaviourist input as habitat degradation continues (after Beissinger [11]).<br />

Box 4. Etho-conservation tackles global warming<br />

The weather systems of the Earth are expected to become more<br />

extreme, perhaps suddenly and with great spatial heterogeneity,<br />

owing to the atmospheric retention of heat energy from anthropogenic<br />

‘greenhouse gas’ production. How animals might react to<br />

climate change is one of the many ‘behavioural unknowns’ caused<br />

by environmental degradation [13] that is being investigated using<br />

Tinbergen’s four questions.<br />

Mechanisms<br />

Cactus-living Drosophila species might respond to climate<br />

change via selection on the timing of their active period in the<br />

circadian clock mechanism [53]. By becoming active at cooler<br />

times of the day, they are able to avoid deleterious exposure to<br />

heat extremes. The physiological stress responses of vertebrate<br />

organisms are likely to be dependent on individual differences in<br />

social status and the social function of dominance in that species<br />

[31,54].<br />

Ontogeny<br />

Culturally determined foraging movements of sperm whale Physeter<br />

macrocephalus clans might make some song clans predictably<br />

susceptible to changes in ocean-current-dependent food sources<br />

[55]. For species with temperature-mediated sex determination,<br />

such as turtles and crocodilians, behavioural commitment to<br />

reusing traditional nest sites will impede adaptive population-level<br />

responses in these long-lived animals. For example, if global<br />

warming occurs as forecast, modelling suggests that natal site<br />

philopatry by egg-laying painted turtles will condemn them to<br />

producing such biased sex ratios that the population becomes<br />

inviable [43].<br />

Function<br />

Species with poor mobility might experience rapid evolutionary<br />

behavioural adaptation in response to micro-climatic divergence.<br />

For example, wood frog tadpoles in shaded, cool ponds appear to<br />

evolve adaptive thermal preferences quickly [56]. Populations in<br />

warmer sun-lit ponds, however, seem to lack growth-benefiting<br />

thermal preferences, suggesting that global warming would cause<br />

meta-population sinks with predictable localized patterns of species<br />

extirpation. Hawksbill turtle Eretmochelys imbricata females might<br />

be able to adjust to a hotter climate through existing preferences for<br />

tree-shaded beach nesting sites. Unfortunately, beach deforestation<br />

may prevent this endangered species from avoiding climateinduced<br />

biased sex ratios [57].<br />

Phylogeny<br />

Sexual selection may have an impact on the response of animals<br />

to climate change. The degree of advancement of spring migration<br />

in birds is associated with the strength of female choice across<br />

species [58]. Colonizing warmer habitats appears to release<br />

energetic constraints on sexual selection in dark-eyed juncos Junco<br />

hyemalis [59]. The relaxation of cold climate extremes might select<br />

against isolating mechanisms, such as reproductive diapause in<br />

mustelids [60].<br />

www.sciencedirect.com<br />

and management to maximize post-release survival of<br />

reintroduced individuals.<br />

Often, captive conditions impose different selection<br />

pressures on animal genomes than natural selection,<br />

resulting in behaviour that is advantageous to the survival<br />

and reproduction of individuals in captivity, but maladaptive<br />

should they be reintroduced to the wild. For example,<br />

captivity selects for aggressive behaviour in place of foraging<br />

behaviour in breeding colonies of the endangered<br />

butterfly splitfin fish Ameca splendens [20]. Behaviourists<br />

are helping to modify aquaculture programs to produce fish<br />

that will forage rather than fight when released to their<br />

restored habitat.<br />

In other cases, the protected nature of captive<br />

environments might allow genetic release from directional<br />

selection. The escape behaviour of captive-bred oldfield<br />

mice Peromyscus polionotus, for example, is never subjected<br />

to selection by owls, stoats, snakes or any of the<br />

predators that normally threaten the survival of free-living<br />

rodents. As a result, they have long escape latencies that<br />

would make them highly susceptible to predators in the<br />

wild [21]. By considering the genetic mechanisms underlying<br />

variation in behaviour, McPhee and Silverman [22]<br />

conclude that we need not abandon reintroduction of such<br />

individuals. Their solution is to simply use the variance in<br />

escape behaviour to recalculate the number of released<br />

animals sufficient to ensure a surviving nucleus of breeders.<br />

By understanding the genetic mechanisms underlying<br />

behaviour, conservation behaviourists are able to<br />

provide release and survival estimates that are more<br />

realistic. This would thus engender more reasonable expectations<br />

and cost planning by wildlife managers and<br />

political decision-makers, and more patience for success<br />

from the general public.<br />

Similarly careful behavioural preparation of captiveraised<br />

animals will lessen animal welfare concerns over<br />

the poor survival of individuals released to the wild for<br />

conservation purposes. Anti-predator training of captivereared<br />

prairie dogs Cynomys ludovicianus appears to<br />

increase survival upon reintroduction [23]. But the value<br />

of behavioural manipulation is not limited to potential prey<br />

species. Poor attention to the social management of African<br />

wild dog Lycaon pictus groups during ‘soft’ releases to the<br />

wild might explain several costly reintroduction failures<br />

[24]. If individuals are chosen to minimize dominance<br />

conflicts among members of the released wild dog pack,<br />

the reintroduced animals are more likely to behave cooperatively<br />

and hopefully survive to reproduce.


Conservation behaviour and natural populations<br />

There are two major areas of inquiry in which ‘behaviour<br />

thinking’ is already being applied to the conservation of<br />

wild populations: improving population viability by adaptively<br />

managing individual survival and reproductive success<br />

in isolated populations, and identifying and managing<br />

deleterious effects of ecotourism.<br />

Managing survival and reproduction<br />

Re-establishing prairie dog populations in protected areas<br />

is important to ecosystem functioning [25], and is crucial to<br />

efforts to establish a prey base to grow viable populations<br />

of the highly endangered black-footed ferret Mustela<br />

nigripes. Behaviourist Debra Shier took advantage of existing<br />

behavioural information on the adaptive nature of kin<br />

groups to demonstrate experimentally that translocating<br />

entire wild prairie dog families achieves conservation goals<br />

more successfully and efficiently than moving unrelated<br />

animals [26]. The red-cockaded woodpecker Picoides borealis<br />

is another endangered US species that has benefited<br />

from a behaviourist’s understanding of kin recognition<br />

mechanisms. Wallace and I showed that, by exchanging<br />

nestlings between nesting cavities at an age young enough<br />

that parents had not yet learned to identify their offspring,<br />

we could overcome genetic isolation in a fragmented<br />

habitat without reducing survivorship of translocated<br />

individuals [27].<br />

Behaviourists have shown how evolutionary conflicts<br />

among individual animals can give us insight into management<br />

methods that would not be apparent to a<br />

traditionally trained wildlife ecologist or conservation<br />

geneticist. When it comes to population viability, larger<br />

population size is usually better. Unfortunately, small<br />

populations do not always have room to grow. A case in<br />

point is the threatened Cuban iguana Cyclura nubile<br />

population confined to the US Naval base at Guantanamo<br />

Bay, Cuba. In the absence of opportunities to<br />

increase the overall lizard population, Alberts et al.<br />

[28] showed how the negative impact of male dominance<br />

on population-wide genetic variation (N e) could be overcome<br />

through behavioural management. Temporarily<br />

removing dominant males allowed other adult males to<br />

obtain mates.<br />

Sustainable ecotourism<br />

Ecotourism, the other doyen of sustainable-use advocates,<br />

must also consider the behaviourally mediated impact of<br />

human disturbance on population viability. Descriptive<br />

behavioural studies are likely to find that wildlife change<br />

their behaviour in response to human visitors, but that<br />

changes in behaviour are not necessarily bad for the<br />

animal under observation. For example, although foraging<br />

brown bears Ursus arctos alter their feeding activities so<br />

that they can be vigilant in the presence of tourists, this<br />

behavioural change has no apparent effects on body condition<br />

[29]. Therefore, human disturbance probably does<br />

not translate into reduced survival and lower population<br />

viability in this case. Nesting Magellanic penguins Spheniscus<br />

magellanicus appear to habituate to frequent tourists,<br />

but previously undisturbed colonies are markedly<br />

stressed by the arrival of human visitors [30]. Because<br />

www.sciencedirect.com<br />

Opinion TRENDS in Ecology and Evolution Vol.22 No.8 405<br />

stressed animals are likely to experience tradeoffs in<br />

reproductive investment or survival [31], opening new<br />

colonies to tourism is not recommended without careful<br />

determination of the human activities that cause the most<br />

stress to nesting penguins. The complexity of the interspecific<br />

and intraspecific responses of wild animals to<br />

anthropogenic and natural stressors benefits from being<br />

managed in a conservation behaviour framework.<br />

Should conservation behaviour conform to the<br />

emphases of conservation biology?<br />

I think the only way that behavioural biology makes<br />

sense for conservation is if we retain our unique<br />

perspective on animal diversity. The ecologists that<br />

helped found the Society for Conservation Biology [32]<br />

did not abandon island biogeography theory to work in<br />

conservation; they applied the concept to the habitat<br />

islands that are nature reserves. Likewise, population<br />

geneticists did not ignore Hardy<strong>–</strong>Weinberg equilibrium<br />

to promote the conservation of bottlenecked populations;<br />

rather, they found ways to inform us of the practical<br />

importance of theoretical genetics. Conservation behaviourists<br />

are conservation biologists; we should be<br />

constructively critical of endangered species recovery<br />

plans [33]. Decisions are often made with very little or<br />

faulty evidence. For example, decisions to allow habitat<br />

destruction in the dwindling range of the endangered<br />

Florida panther Puma concolor coryi were based on<br />

questionable evidence and the illogical opinion of one<br />

researcher that this subspecies is a forest obligate (for a<br />

shocking review, see [34]).<br />

It has become a conservation platitude of sorts that<br />

conservation behaviourists must ‘‘translate behavior into<br />

currencies relevant to conservation at large spatial scales’’<br />

[11]. But conflicts with large carnivores and other species of<br />

potential harm to humans attract much popular press<br />

and political interest. It is the young panther that<br />

disperses 350 miles [34] or the few African elephants<br />

Loxodonta africana that invade villages or kill rhinos<br />

[35,36] that threaten to scuttle goodwill for conservation,<br />

not the average behaviour of the population. It is ridiculous<br />

to suggest that individualistic responses of animals are<br />

unimportant to conservation.<br />

Conservation behaviourists will rise to the challenge<br />

of the important discoveries being made at the interface<br />

of conservation ecology and genetics. For example, Riley<br />

and colleagues [37] recently used microsatellite analysis<br />

combined with radiotelemetry to discover that carnivore<br />

territories tend to pile up at habitat edges along an<br />

automotive freeway in California, USA. The intense<br />

social challenges faced by individual bobcats Lynx rufus<br />

and coyotes Canis latrans attempting to disperse<br />

through the gauntlet of territories concentrated near<br />

large roadways accentuated the limits on gene flow<br />

imposed by the physical barrier of the road itself. Populations<br />

on either side of the highway show evidence of<br />

genetic differentiation as a result. The problem of<br />

‘territory piling’ [37] along roadways is one well suited<br />

to Tinbergen’s framework. These crowded carnivores<br />

should expect a visit from some local conservation<br />

behaviourists [38].


406 Opinion TRENDS in Ecology and Evolution Vol.22 No.8<br />

The next step<br />

Caro’s [3] charges of irrelevancy levied against conservation<br />

behaviour are genuine, but not unique to our field.<br />

There are problems with the application of any science to<br />

public policy. Academic scientists get caught up in grant<br />

writing, hypotheses testing and data collection, whereas<br />

conservation practitioners need practical advice today [39].<br />

The reverse is also true; conservation monitoring programs<br />

often take on a life of their own and do not achieve conservation<br />

objectives [40].<br />

Nearly 100 years ago, conservationist William T Hornaday<br />

[41] declared: ‘‘We will endeavor to avoid the discussion<br />

of academic questions, because the business of<br />

conservation is replete with urgent practical demands’’. I<br />

believe that it is this sort of concern, that there is an<br />

opportunity cost to the ‘theoretical’ concerns of conservation<br />

behaviour, that has led some of my colleagues to<br />

retreat from conservation behaviour. Instead, they must<br />

come to realize that we approach a time of desperate need<br />

for applied behaviourists. One need only look at costly<br />

efforts to recover species listed under the US Endangered<br />

Species Act [42] to see that, although all is not lost, all is not<br />

well either. Wildlife managers have stopped many endangered<br />

species from going extinct, but few species have<br />

recovered sufficiently to be de-listed. We can improve<br />

conservation management decisions. The behaviour of<br />

individual animals matters to conservation. The growing<br />

pains of conservation behaviour are not symptoms of dysfunction,<br />

but rather positive signs of a thriving adolescence.<br />

Conservation behaviour is relevant right now and<br />

the time is ripe for the conservation behaviourist to make a<br />

difference.<br />

Acknowledgements<br />

I am grateful to Cliff Ochs, Guillermo Paz y Miño and Suhel Quader for<br />

commenting on an earlier draft of this paper. Detailed suggestions by Dan<br />

Blumstein, Marco Festa-Bianchet and an anonymous reviewer improved<br />

the manuscript. Communications with Steve Beissinger and Joel Berger<br />

were helpful ten years ago and now too. Tim Caro suggested that I put it<br />

in writing, and I am very appreciative of his and the editors’ assistance<br />

and generous patience. Jill Frank and Glenn Parsons kindly offered the<br />

use of their dolphin photos.<br />

References<br />

1 Soulé, M.E. (1985) What is conservation biology? Bioscience 35, 727<strong>–</strong><br />

734<br />

2 Buchholz, R. (2006) Should animal behaviorists teach conservation?<br />

Conserv. Behav. 4, 3<strong>–</strong>4<br />

3 Caro, T. (2007) Behaviour and conservation: a bridge too far? Trends<br />

Ecol. Evol. (in press)<br />

4 Knight, J. (2001) If they could talk to the animals.... Nature 414, 246<strong>–</strong><br />

247<br />

5 Clemmons, J.R. and Buchholz, R., eds (1997) Behavioral Approaches<br />

to Conservation in the Wild, Cambridge University Press<br />

6 Arcese, P. et al. (1997) Why hire a behaviorist into a conservation or<br />

management team? In Behavioral Approaches to Conservation in the<br />

Wild (Clemmons, J.R. and Buchholz, R., eds), pp. 48<strong>–</strong>71, Cambridge<br />

University Press<br />

7 Greene, H.W. (2005) Organisms in nature as a central focus for biology.<br />

Trends Ecol. Evol. 20, 23<strong>–</strong>27<br />

8 Tinbergen, N. (1963) On aims and methods of ethology. Z. Tierpsychol.<br />

20, 410<strong>–</strong>433<br />

9 Blumstein, D.T. (2007) Tinbergen’s four questions. In The Encyclopedia<br />

of Applied Animal Behaviour and Welfare (Mills, D. ed), (in press), CAB<br />

<strong>International</strong><br />

www.sciencedirect.com<br />

10 Schlaepfer, M.A. et al. (2002) Ecological and evolutionary traps. Trends<br />

Ecol. Evol. <strong>17</strong>, 474<strong>–</strong>480<br />

11 Beissinger, S.R. (1997) Integrating behavior into conservation biology:<br />

potentials and limitations. In Behavioral Approaches to Conservation<br />

in the Wild (Clemmons, J.R. and Buchholz, R., eds), pp. 23<strong>–</strong>47,<br />

Cambridge University Press<br />

12 Sanderson, E.W. et al. (2002) The human footprint and the last of the<br />

wild. Bioscience 52, 891<strong>–</strong>904<br />

13 Paz-y-Miño, C. (2006) Behavioral unknowns: an emerging challenge for<br />

conservation. Conserv. Behav. 4, 2<br />

14 Myers, N. (1995) Environmental unknowns. Science 269, 358<strong>–</strong>360<br />

15 Fortin, D.L. et al. (2005) Wolves influence elk movements: behavior<br />

shapes a trophic cascade in Yellowstone National Park. Ecology 86,<br />

1320<strong>–</strong>1330<br />

16 Berger, J. Carnivore repatriation and Holarctic prey: narrowing the<br />

deficit in ecological effectiveness. Conserv. Biol. (in press)<br />

<strong>17</strong> Purvis, A. et al. (2000) Predicting extinction risk in declining species.<br />

Proc. Biol. Sci. 267, 1947<strong>–</strong>1952<br />

18 Sutherland, W.J. (2006) Predicting the ecological consequences of<br />

environmental change: a review of the methods. J. Appl. Ecol. 43,<br />

599<strong>–</strong>616<br />

19 Snyder, N.F.R. et al. (1996) Limitations of captive breeding in<br />

endangered species recovery. Conserv. Biol. 10, 338<strong>–</strong>348<br />

20 Kelley, J.L. (2006) Captive breeding promotes aggression in an<br />

endangered Mexican fish. Biol. Conserv. 133, 169<strong>–</strong><strong>17</strong>7<br />

21 McPhee, M.E. (2004) Generations in captivity increases behavioral<br />

variance: considerations for captive breeding and reintroduction<br />

programs. Biol. Conserv. 115, 71<strong>–</strong>77<br />

22 McPhee, M.E. and Silverman, E.D. (2004) Increased behavioral<br />

variation and the calculation of release numbers for reintroduction<br />

programs. Conserv. Biol. 18, 705<strong>–</strong>715<br />

23 Shier, D.M. and Owings, D.H. (2006) Effects of predator training on<br />

behavior and post-release survival of captive prairie dogs (Cynomys<br />

ludovicianus). Biol. Conserv. 132, 126<strong>–</strong>135<br />

24 Gusset, M. et al. (2006) Divided we fail: the importance of social<br />

integration for the re-introduction of endangered African wild dogs<br />

(Lycaon pictus). J. Zool. 270, 502<strong>–</strong>511<br />

25 Davidson, A.D. and Lightfoot, D.C. (2006) Keystone rodent<br />

interactions: prairie dogs and kangaroo rats structure the biotic<br />

composition of a desertified grassland. Ecography 29, 755<strong>–</strong>765<br />

26 Shier, D.M. (2006) Effect of family support on the success of<br />

translocated blacktailed prairie dogs. Conserv. Biol. 20, <strong>17</strong>80<strong>–</strong><strong>17</strong>90<br />

27 Wallace, M.T. and Buchholz, R. (2001) Translocation of red-cockaded<br />

woodpeckers by reciprocal fostering of nestlings. J. Wildl. Manage. 65,<br />

327<strong>–</strong>333<br />

28 Alberts, A.C. et al. (2003) Temporary alteration of local social structure<br />

in a threatened population of Cuban iguanas (Cyclura nubila). Behav.<br />

Ecol. Sociobiol. 51, 324<strong>–</strong>335<br />

29 Rode, K.D. et al. (2006) Behavioural responses of brown bears mediate<br />

nutritional effects of experimentally introduced tourism. Biol. Conserv.<br />

133, 70<strong>–</strong>80<br />

30 Walker, B.G. et al. (2006) Habituation of adult Magellanic penguins to<br />

human visitation as expressed through behaviour and corticosterone<br />

secretion. Conserv. Biol. 20, 146<strong>–</strong>154<br />

31 Wingfield, J.C. and Sapolsky, R.M. (2003) Reproduction and resistance<br />

to stress: when and how. J. Neuroendocrinol. 15, 711<strong>–</strong>724<br />

32 Ralls, K. (1997) On becoming a conservation biologist. In Behavioral<br />

Approaches to Conservation in the Wild (Clemmons, J.R. and Buchholz,<br />

R., eds), pp. 356<strong>–</strong>372, Cambridge University Press<br />

33 Boersma, P.D. et al. (2001) How good are endangered species recovery<br />

plans? Bioscience 51, 643<strong>–</strong>649<br />

34 Gross, L. (2005) Why not the best? How science failed the Florida<br />

panther. PLoS Biol. 3, e333<br />

35 Rode, K.D. et al. (2006) Nutritional ecology of elephants in Kibale<br />

National Park, Uganda, and its relationship with crop-raiding<br />

behaviour. J. Trop. Ecol. 22, 441<strong>–</strong>449<br />

36 Slowtow, R. and van Dyk, G. (2001) Role of delinquent young ‘‘orphan’’<br />

male elephants in high mortality of white rhinoceros in Pilanesberg<br />

National Park, South Africa. Koedoe 44, 85<strong>–</strong>94<br />

37 Riley, S.P.D. et al. (2006) A southern California freeway is a physical<br />

and social barrier to gene flow in carnivores. Mol. Ecol. 15, <strong>17</strong>33<strong>–</strong><strong>17</strong>41<br />

38 Blumstein, D.T. and Fernández-Juricic, E. (2004) The emergence of<br />

conservation behavior. Conserv. Biol. 18, 1<strong>17</strong>5<strong>–</strong>1<strong>17</strong>7


39 Cabin, R.J. (2007) Science-driven restoration: a square grid on a round<br />

Earth? Restor. Ecol. 15, 1<strong>–</strong>7<br />

40 Nichols, J.D. and Williams, B.K. (2006) Monitoring for conservation.<br />

Trends Ecol. Evol. 21, 668<strong>–</strong>673<br />

41 Hornaday, W.T. (1914) Wild Life Conservation in Theory and Practice,<br />

Yale University Press<br />

42 Male, T.D. and Bean, M.J. (2005) Measuring progress in US<br />

endangered species conservation. Ecol. Lett. 8, 986<strong>–</strong>992<br />

43 Morjan, C. (2003) How rapidly can maternal behavior affecting<br />

primary sex ratio evolve in a reptile with environmental sex<br />

determination? Am. Nat. 162, 205<strong>–</strong>219<br />

44 Caro, T. ed. (1998) Behavioural Ecology and Conservation Biology,<br />

Oxford University Press<br />

45 Linklater, W.L. (2004) Wanted for conservation research:<br />

behavioral ecologists with a broader perspective. Bioscience 54,<br />

352<strong>–</strong>360<br />

46 Soulé, M.E. (1986) Conservation biology and the ‘‘real world’’. In<br />

Conservation Biology: the Science of Scarcity and Diversity (Soulé,<br />

M.E., ed.), pp. 1<strong>–</strong>12, Sinauer Associates<br />

47 Read, A.J. et al. (1993) Harbor porpoises and gill nets in the Gulf of<br />

Maine. Conserv. Biol. 7, 189<strong>–</strong>193<br />

48 Carlstrom, J. et al. (2002) A field experiment using acoustic alarms<br />

(pingers) to reduce harbour porpoise by-catch in bottom-set gillnets.<br />

ICES J. Mar. Sci. 59, 816<strong>–</strong>824<br />

49 Larsen, F. et al. (2002) Reduction in harbour porpoise by-catch in the<br />

North Sea by high density gillnets. Paper SC/54/SM30, <strong>International</strong><br />

Whaling Commission Scientific Committee (www.cetaceanbycatch.org/<br />

Papers/larse02.pdf)<br />

A Special Issue of Current Biology<br />

21st August 2007<br />

50 Cox, T.M. et al. (2003) Behavioral responses of bottlenose dolphins,<br />

Tursiops truncatus, to gillnets and acoustic alarms. Biol. Conserv. 115,<br />

203<strong>–</strong>212<br />

51 Kastelein, R.A. et al. (2006) Differences in the response of a striped<br />

dolphin (Stenella coeruleoalba) and a harbour porpoise (Phocoena<br />

phocoena) to an acoustic alarm. Mar. Environ. Res. 61, 363<strong>–</strong>378<br />

52 Naidoo, R. et al. (2006) Integrating economic costs into conservation<br />

planning. Trends Ecol. Evol. 21, 681<strong>–</strong>687<br />

53 Sørensen, G.K. and Loeschcke, V. (2002) Natural adaptation to<br />

environmental stress via physiological clock-regulation of stress<br />

resistance in Drosophila. Ecol. Lett. 5, 16<strong>–</strong>19<br />

54 Cockrem, J.F. (2005) Conservation and behavioral neuroendocrinology.<br />

Horm. Behav. 48, 492<strong>–</strong>501<br />

55 Whitehead, H. and Rendell, L. (2004) Movements, habitat use and<br />

feeding success of cultural clans of South Pacific sperm whales.<br />

J. Anim. Ecol. 73, 190<strong>–</strong>196<br />

56 Freidenburg, L.K. and Skelly, D.K. (2004) Microgeographical variation<br />

in thermal preference by an amphibian. Ecol. Lett. 7, 369<strong>–</strong>373<br />

57 Kamel, S.J. and Mrosovsky, N. (2006) Deforestation: risk of sex ratio<br />

distortion in Hawksbill sea turtles. Ecol. Appl. 16, 923<strong>–</strong>931<br />

58 Spottiswoode, C.N. et al. (2006) Sexual selection predicts advancement<br />

of avian migration in response to climate change. Proc. R. Soc. Lond. B.<br />

Biol. Sci. 273, 3023<strong>–</strong>3029<br />

59 Yeh, P.J. (2004) Rapid evolution of a sexually selected trait following<br />

population establishment in a novel habitat. Evolution 58, 166<strong>–</strong><strong>17</strong>4<br />

60 Thom, M.D. et al. (2004) The evolution and maintenance of delayed<br />

implantation in the Mustelidae (Mammalia: Carnivora). Evolution 58,<br />

<strong>17</strong>5<strong>–</strong>183<br />

This special issue of Current Biology takes a broad look at the importance of sociality in biology, from the evolution of<br />

cooperation to the way that social living might have facilitated the evolution of human intelligence. The issue<br />

illustrates the fascinating variety of biological societies with articles on social amoebae, social spiders, eusocial<br />

insets, crows, hyenas and humans.<br />

Guest Editorial<br />

All life is social<br />

Steve Frank<br />

Reviews<br />

Evolutionary explanations for cooperation<br />

Stuart West, Ashleigh Griffin and Andy Gardner<br />

Kin selection versus sexual selection in eusocial<br />

insects<br />

Jacobus Boomsma<br />

Social learning<br />

Lars Chittka<br />

The Cold War of the social amoebae<br />

Gad Shaulsky and Richard Kessin<br />

Social cognition in humans<br />

Chris Frith<br />

Sociality, evolution and cognition<br />

Richard Byrne and Lucy Bates<br />

www.sciencedirect.com<br />

Opinion TRENDS in Ecology and Evolution Vol.22 No.8 407<br />

Social immune systems<br />

Sylvia Cremer<br />

Primers<br />

The social life of corvids<br />

Nicky Clayton and Nathan Emery<br />

Hyena societies<br />

Heather Watts and Kay Holekamp<br />

Social spiders<br />

Duncan Jackson<br />

Quick Guide


Anim Cogn (2009) 12:43<strong>–</strong>53<br />

DOI 10.1007/s10071-008-0169-9<br />

ORIGINAL PAPER<br />

Evidence of teaching in atlantic spotted dolphins<br />

(Stenella frontalis) by mother dolphins foraging in the presence<br />

of their calves<br />

Courtney E. Bender Æ Denise L. Herzing Æ<br />

David F. Bjorklund<br />

Received: 10 January 2008 / Revised: 18 March 2008 / Accepted: 5 June 2008 / Published online: 29 July 2008<br />

Ó Springer-Verlag 2008<br />

Abstract Teaching is a powerful form of social learning,<br />

but there is little systematic evidence that it occurs in<br />

species other than humans. Using long-term video archives<br />

the foraging behaviors by mother Atlantic spotted dolphins<br />

(Stenella frontalis) were observed when their calves were<br />

present and when their calves were not present, including<br />

in the presence of non-calf conspecifics. The nine mothers<br />

we observed chased prey significantly longer and made<br />

significantly more referential body-orienting movements in<br />

the direction of the prey during foraging events when their<br />

calves were present than when their calves were not present,<br />

regardless of whether they were foraging alone or with<br />

another non-calf dolphin. Although further research into<br />

the potential consequences for the naïve calves is still<br />

warranted, these data based on the maternal foraging<br />

behavior are suggestive of teaching as a social-learning<br />

mechanism in nonhuman animals.<br />

Keywords Teaching Dolphins Social learning<br />

Foraging<br />

Electronic supplementary material The online version of this<br />

article (doi:10.1007/s10071-008-0169-9) contains supplementary<br />

material, which is available to authorized users.<br />

C. E. Bender (&) D. F. Bjorklund<br />

Department of Biological Sciences, Florida Atlantic University,<br />

777 Glades Road, Boca Raton, FL 33431-0991, USA<br />

e-mail: courtbender@yahoo.com; cgreen21@fau.edu<br />

D. L. Herzing<br />

Department of Biological Sciences, Wild Dolphin Project,<br />

Florida Atlantic University, 777 Glades Road, Boca Raton,<br />

FL 33431-0991, USA<br />

Introduction<br />

Although one of the most potent forms of social learning in<br />

humans, there has been little evidence to suggest that<br />

teaching occurs in nonhuman animals. Some theorists have<br />

suggested that this may be because teaching requires<br />

advanced social-cognitive skills, including the ability to<br />

take the perspective of another and theory of mind, the<br />

ability to appreciate that an individual’s behavior is based<br />

on its knowledge and its desires (Boesch and Tomasello<br />

1998; Tomasello 1996, 2000; Tomasello et al. 1993). For<br />

example, previous examples of suggested teaching behavior<br />

by meerkats, cheetahs, and domestic cats seem to<br />

benefit the prey-handling abilities of the young, but do not<br />

require the use of higher cognitive mechanisms (Thornton<br />

and McAuliffe 2006; Caro and Hauser 1992). However,<br />

Caro and Hauser (1992) provided a definition of teaching<br />

that may be more inclusive for nonhuman animals, defining<br />

it as, ‘‘An individual actor A can be said to teach if it<br />

modifies its behavior only in the presence of a naive<br />

observer, B, at some cost or at least without obtaining an<br />

immediate benefit for itself. A’s behavior thereby encourages<br />

or punishes B’s behavior, or provides B with<br />

experience, or sets an example for B. As a result, B<br />

acquires knowledge or learns a skill earlier in life or more<br />

rapidly or efficiently than it might otherwise do, or that it<br />

would not learn at all’’ (p. 153).<br />

Previous studies that suggested teaching in primates and<br />

cetaceans, although promising, lacked systematic measurement<br />

of the behavior. Probably the best evidence of<br />

teaching in nonhuman primates to date is Boesch’s studies<br />

of mother chimpanzees (Pan troglodytes) in the Tai<br />

National Park of the Ivory Coast (Boesch 1991, 1993;<br />

Greenfield et al. 2000). Boesch suggested that the<br />

chimpanzee mothers facilitated the development of their<br />

123


44 Anim Cogn (2009) 12:43<strong>–</strong>53<br />

offspring’s nut-cracking skills by means of stimulation,<br />

facilitation, and active teaching. Nut cracking is observed<br />

in only a few populations of chimpanzees, despite the<br />

availability of nuts and appropriate tools (i.e., rocks<br />

appropriate for use as anvils and hammers), qualifying, by<br />

some definitions, as an example of culturally-transmitted<br />

behavior (Whiten 2005; Whiten et al. 1999). The interpretation<br />

of such episodes as ‘‘teaching’’ has been<br />

questioned, however (Bering 2001; Bering and Povinelli<br />

2003). Moreover, such episodes are rarely observed, suggesting<br />

that direct teaching is not a common form of<br />

cultural transmission in chimpanzees.<br />

Although evidence of social learning is easier to document<br />

in these terrestrial great apes than it is in marine<br />

mammals, nongenetic transmission of behavior across<br />

generations has also been observed for cetaceans (Kruetzen<br />

et al. 2005; Rendell and Whitehead 2001), suggesting that,<br />

like the great apes, these large-brained, slow-developing,<br />

and socially complex species (Bjorklund and Bering 2003)<br />

have evolved powerful social-learning mechanisms.<br />

Similar to research with great apes, little is known about<br />

the actual mechanism of transmission across generations in<br />

cetaceans. Herzing (2005) described some potential scenarios<br />

and mechanisms observed for a group of free-ranging<br />

Atlantic spotted dolphins (Stenella frontalis), including<br />

implications of vertical, horizontal, and oblique directions<br />

of transmission of information during various behavioral<br />

contexts. Recently, Spininelli et al. (2006) described preytransfer<br />

between mother and calf in the marine tucuxi dolphin<br />

(Sotalia fluvialis). Observations such as these suggest<br />

that the mother-calf relationship may be one of the most<br />

important sources of information in the young calf’s life.<br />

There is some evidence of presumed maternal teaching<br />

behavior associated with stranding behavior as a foraging<br />

specialization used by part of the population of killer whales<br />

(Orcinus orca) in the Crozet Islands and off Punta Norte,<br />

Argentina to capture seal pups on pinniped breeding beaches.<br />

Adult females demonstrated a modification of their strand<br />

foraging behavior in the presence of naïve juvenile observers<br />

(presumably their calves), suggesting that teaching may be<br />

involved in the development and the rate of success of calves<br />

in mastering these behaviors (Guinet and Bouvier 1995).<br />

This comparison between purported teaching in chimpanzees<br />

and killer whales is interesting because any<br />

commonalities would have been derived through convergent<br />

evolution, as the last common ancestor of primates with<br />

cetaceans is estimated to have lived over 90 million years ago<br />

(Marino et al. 2007). However, despite the multitude of<br />

observations of chimpanzees and killer whales in the wild,<br />

incidences of mother<strong>–</strong>infant teaching are scarce and<br />

anecdotal in nature.<br />

Foraging behavior is a likely candidate for social<br />

learning among wild dolphins, particularly between mother<br />

123<br />

and calf. The mother/calf relationship is the strongest<br />

association that Atlantic spotted dolphins have in their<br />

lifetimes (Herzing and Brunnick 1997). The prolonged<br />

developmental period provides both ample time and situational<br />

possibilities for a calf to learn foraging strategies<br />

from its mother (Herzing 1996). The majority of daytime<br />

feeding behavior of Atlantic spotted dolphins is benthic<br />

foraging, in which dolphins use echolocation to locate fish<br />

in the sandy bottom and then dig prey items out of the sand<br />

in order to catch and eat them (Herzing 1996).<br />

Much of what we know about marine mammal social<br />

learning comes from research with captive animals due, in<br />

part, to the difficulty of studying such phenomena in the<br />

wild. Attempts to assess teaching among captive animals<br />

often involve contrived situations, which may affect the<br />

animals’ success or failure (Kuczaj et al. 2005). Studies<br />

that assess social learning in wild populations of marine<br />

mammals are needed to validate and supplement the findings<br />

from captive animals and to better understand the<br />

spontaneous occurrence of social learning in natural<br />

settings.<br />

In the present study, using video archives from a longterm<br />

naturalistic study, we investigated social learning and<br />

possible teaching behavior by the Atlantic spotted dolphin.<br />

Unlike most previous research examining social learning in<br />

nonhuman mammals in which the focus is on the observer/<br />

learner (e.g., Bjorklund et al. 2002; Guinet and Bouvier<br />

1995; Herman 2002; Tomasello et al. 1993), the present<br />

study shifted the focus from the observer (in our case, the<br />

calf observing the foraging behavior) to the model (the<br />

mother performing the foraging behavior) to explore the<br />

possibility of teaching behavior. We examined the foraging<br />

behavior by mother dolphins when foraging in the presence<br />

of their young (less than 3-year-old) calves in comparison<br />

to when the calves were not present. Additionally, comparisons<br />

were then made between mothers foraging alone<br />

and when they were foraging with non-calf conspecifics.<br />

Should the mothers alter their foraging behavior in the<br />

presence of their young, it would be suggestive of teaching<br />

and provide a possible mechanism for cultural transmission<br />

within this dolphin species.<br />

Methods and analysis<br />

Natural history<br />

This study was performed using underwater video recordings<br />

of the Atlantic spotted dolphin collected by the Wild<br />

Dolphin Project (Herzing 1997) in the study area north of<br />

Grand Bahama Island, Bahamas during summer field seasons<br />

between 1991 and 2004 (Fig. 1). Unlike many other<br />

marine mammal habitats, this area is optimal for behavioral


Anim Cogn (2009) 12:43<strong>–</strong>53 45<br />

Fig. 1 Map of study area for the wild dolphin project underwater<br />

video recordings collected over sandbanks north of Grand Bahama<br />

Island Study population of Atlantic spotted dolphins ranges over an<br />

area of approximately 500 km 2<br />

observation with clear, warm waters that allow for excellent<br />

visibility up to 90 ft and long observational periods<br />

(Herzing 1996). The dolphins in this study have been<br />

observed since 1985, and include over 200 dolphins that<br />

have been individually recognized and sexed. Atlantic<br />

spotted dolphins can be categorized into four age classes,<br />

based on the pigmentation of the individual (Herzing<br />

1996). The number of spots on the individual is correlated<br />

with age, with a newborn having no spots. Although most<br />

individuals in the population are tracked from birth, this<br />

allows for approximating the individual’s age when it is not<br />

known from previous sightings. Calves in the sampled<br />

foraging events 3-years-old or younger were still observed<br />

to be nursing during the encounter year. The year of the<br />

calf’s birth was determined from previous sightings of the<br />

pregnant female followed by a sighting with a closely<br />

associated calf or a sighting of the mother with a suckling<br />

calf (Herzing 1997).<br />

Apparatus and procedure<br />

Video was recorded using various underwater cameras<br />

(Sony CCDV9 8-mm, Yashica KXV Hi8-mm with attached<br />

Labcore 76 hydrophone, Sony DCR-SC100 NTSC, or Sony<br />

DCR-PC110 NTSC Digital Video). Underwater video<br />

sequences were analyzed using focal follow (of the mother<br />

with and without the calf) as a sampling rule and continuous<br />

sampling as a recording rule. Video sequences from<br />

the long-term video archives from the Wild Dolphin Project<br />

Ò between 1991 and 2004 were assessed for the<br />

presence of benthic foraging events by individual mothers,<br />

either with or without their calves present, based on a<br />

behavioral ethogram designed to measure the individual<br />

benthic foraging behavior of the mother dolphins (see<br />

Fig. 2). Each benthic fish catch was broken down into a<br />

series of behaviors that made up a typical foraging event.<br />

For the purpose of this study, a foraging event was<br />

defined as the series of behaviors performed by the dolphin<br />

in order to catch the prey animal. For benthic feeding, the<br />

series of events was as follows:<br />

scanning ! rooting ! chase ! ingestion<br />

Scanning is observed when the dolphin moves its head<br />

horizontally or vertically repeatedly while performing a<br />

directional swim, and usually occurs near the sea floor and<br />

can be followed by a dig or fish catch (Miles and Herzing<br />

2003). Rooting or digging is observed when the dolphin<br />

inserts the rostrum into the sea floor or sandy bottom to dig<br />

the prey out of the substrate in the attempt to capture the<br />

prey. Chasing is observed as swimming in the direction of<br />

the prey object, as in pursuit of the prey. For our purposes,<br />

ingestion was defined as the food going into mouth and<br />

never being seen again. The categories of foraging<br />

behaviors measured were chase latency and number of<br />

body-orienting movements during pursuit. Chase latency,<br />

the length of time the prey was pursued, was operationally<br />

defined as the period of time when the fish appeared out of<br />

the sand (was rooted out) until the time of ingestion by the<br />

dolphin, or the dolphin no longer pursued prey (lost interest<br />

in prey). Body-orienting movements were measured to<br />

examine the dolphin’s attention to the prey object. A bodyorienting<br />

movement was measured in foraging events as a<br />

movement of the body reorienting in direction of prey<br />

object, often seen during pursuit of prey, from the time the<br />

prey was rooted out of the sand. Body-orienting<br />

movements were particularly interesting as they appeared<br />

to be exaggerated movements in the direction of the prey,<br />

which may be an attention-directed referential behavior<br />

similar to the spontaneous pointing observed by dolphins<br />

during experiments in captivity (Xitco et al. 2001).<br />

Figure 2 is a visual ethogram describing the sequence of<br />

the benthic foraging behavior (Miles and Herzing 2003).<br />

123


46 Anim Cogn (2009) 12:43<strong>–</strong>53<br />

Fig. 2 Visual ethogram of<br />

select foraging behavioral<br />

events (Miles and Herzing<br />

2003). The visual ethogram<br />

includes the possible positions<br />

in which calves were observed<br />

during foraging events in<br />

relation to their mother and the<br />

sequence of behaviors for a<br />

benthic foraging event<br />

The segments from the video archives in which mothers<br />

were observed engaged in foraging behavior with or<br />

without their calves were then viewed and then examined<br />

for further criteria. Thirty-eight video segments were used<br />

in the study based on the selection criteria: fourteen video<br />

segments of mothers foraging with their calves present and<br />

24 video segments of those mothers foraging without their<br />

calves present. Videos were selected based on the presence<br />

of a target female performing an individual foraging event,<br />

as well as on the following designated video acceptance<br />

criteria: (a) the individuals were identifiable; (b) it was<br />

possible to identify the beginning and end of the chase<br />

sequence; (c) the prey was visible, or if the prey was not<br />

completely visible, it was possible to identify the position<br />

of the prey based on the behavior of the dolphin; (d) if the<br />

calf was present, the calf was in a nearby position (within a<br />

proximity of two body lengths) from which it was capable<br />

of observing its mother; and (e) if the calf was present, it<br />

was possible to identify the position of the calf relative to<br />

the mother during the foraging event. The position of the<br />

calf relative to the mother was recorded as Infant, Headunder-head,<br />

Echelon, Observation, or Other, with comments<br />

where necessary, as depicted in Fig. 2.<br />

The mothers’ foraging behaviors were then individually<br />

measured for the variables of chase latency and body-orienting<br />

movements, both when foraging alone (or with other<br />

juvenile or adult dolphins) and when foraging in the<br />

presence of their calves. Other variables, such as types of<br />

play involved in that foraging encounter, position of calf<br />

relative to the mother, directionality of the calf, age of the<br />

mothers and their calves, whether the mother eventually ate<br />

the prey; the prey species, when identifiable, were also<br />

recorded for each foraging event.<br />

Participants<br />

Positions of Calf Relative to Mother<br />

The foraging behaviors of nine mother dolphins were<br />

recorded both with (n = 14) and without (n = 22) their<br />

123<br />

Echelon Position Infant Position Head-under-head Observation Position<br />

Sequence of Behaviors for<br />

Benthic Foraging Event<br />

Position<br />

Individual Scanning Dig/Root Chase<br />

calves present. Ten different calves were observed with the<br />

nine mothers in the 14 foraging events, with one mother<br />

observed during separate events with two different calves.<br />

Calves ranged in age from neonate to 3 years old. All<br />

calves were observed to be nursing within the same field<br />

season as the foraging event. Ages of the mothers were<br />

known, or were estimated based on their age class. It is<br />

important to note that some foraging events without calves<br />

present (10 of the 22) occurred while the target female was<br />

still a juvenile (prior to sexual maturity); however, during<br />

those events, the ‘‘mothers to be’’ were already past the age<br />

of weaning and were independently foraging. The minimum<br />

age of any female during foraging events over the<br />

12-year period was 10 years for mothers foraging with<br />

their calves and 4 years for mothers-to-be foraging without<br />

calves present. Of the observations without calves present,<br />

four of the females were observed foraging as juveniles,<br />

prior to becoming mothers. Of the nine mothers observed,<br />

foraging events were observed for one female both when<br />

she was a calf with her mother, and later as a mother<br />

herself with her own calf.<br />

The video segments of the foraging events were shortened<br />

to within 1 min of the beginning and end of the<br />

foraging event and labeled with the foraging event number<br />

and the individuals involved. Segments were then watched<br />

by the first author and two independent observers to measure<br />

the desired behaviors. Of the 38 total foraging events<br />

measured, 32 were measured by the first author and two<br />

independent observers; the other six were measured by the<br />

first author and only one independent observer. For the<br />

measurement of chase latencies of foraging events, there<br />

was significant correlation between the author and the first<br />

independent observer, r 36 = 1.0, P \ 0.001, between the<br />

author and the second independent observer, r30 = 0.999,<br />

P \ 0.001, and between the first and second observer,<br />

r30 = 0.998, P \ 0.001. For the measurement of number of<br />

body-orienting movements, there was significant correlation<br />

between the author and the first independent observer,


Anim Cogn (2009) 12:43<strong>–</strong>53 47<br />

Table 1 Mean chase latency of mothers with and without calf<br />

Mother Little Gash Mugsy Nassau Nippy PR1 PR2 Rosemole Trimy Uno Mean SD<br />

Mean chase latency<br />

without calf (n)<br />

Mean chase latency<br />

with calf (n)<br />

Mean number of BOM<br />

without calf (n)<br />

Mean number of BOM<br />

with calf (n)<br />

r36 = 1.0, P \ 0.001, between the author and the second<br />

independent observer, r30 = 0.988, P \ 0.001, and<br />

between the first and second observer, r30 = 0.988,<br />

P \ 0.001.<br />

The videos were watched and timed using Windows<br />

Media Player version 10 on a Hewlett Packard laptop<br />

computer and a projector in order to enlarge the viewing<br />

area. Due to the restrictions of the media software used for<br />

editing and playing the video, a ‘‘second or less’’ rule was<br />

instituted for measurement of latencies in which the<br />

latencies appearing to be less than 1 s were rounded up to<br />

1 s in duration.<br />

Results<br />

Chase latencies<br />

Mean chase latencies for each of the nine mothers, both<br />

when foraging with and without their calves, is presented in<br />

Table 1. The mothers chased the prey significantly longer<br />

when their calves were present (M = 22.24 s, SD = 9.36)<br />

than when their calves were not present (M = 2.74 s,<br />

SD = 1.47), t8 = 6.57, P \ 0.001, d = 1.14. Mean chase<br />

latencies were longer when foraging with their calves than<br />

without their calves for each of the nine mothers (Fig. 3).<br />

Body-orienting movements<br />

5.6 (5) 2.25 (4) 2.00 (3) 2.00 (1) 2.<strong>17</strong> (2) 4.00 (1) 1.67 (3) 4.00 (4) 1.00 (1) 2.74 1.47<br />

38.33 (1) 23.00 (2) 16.33 (2) 36.00 (1) 24.33 (3) 19.67 (1) 16.50 (2) 12.00 (1) 14.00 (1) 22.24 9.36<br />

0.80 (5) 0.00 (4) 0.33 (3) 0.67 (1) 0.50 (2) 0.00 (1) 0.00 (3) 0.25 (4) 0.00 (1) 0.28 0.31<br />

0.00 (1) 2.50 (2) 1.50 (2) 0.00 (1) 2.33 (3) 2.00 (1) 0.00 (2) 2.00 (1) 1.00 (1) 1.26 1.04<br />

Bold indicates mothers who performed significantly more body-orienting movements in the presence of their calves. Chase latencies were<br />

significantly longer for all nine mothers<br />

Mean number of body-orienting movements for each of the<br />

nine mothers, both when foraging with and without their<br />

calves, is presented in Table 1. The mothers made significantly<br />

more body-orienting movements when their calves<br />

were present (M = 1.26, SD = 1.04) than when their<br />

calves were not present (M = 0.28, SD = 0.31), t8 = 2.46,<br />

P = 0.04, d = 2.31. Six of nine mothers made more bodyorienting<br />

movements when foraging in the presence of<br />

their calves than when not foraging with their calves.<br />

45.00<br />

40.00<br />

35.00<br />

30.00<br />

25.00<br />

20.00<br />

15.00<br />

10.00<br />

5.00<br />

0.00<br />

38.33<br />

5.60<br />

23.00<br />

Chase Latency of Mothers<br />

16.33<br />

36.00<br />

Cross-generational comparisons<br />

Two consecutive generations of the mother/calf pairs in the<br />

study showed the same altered foraging behavior. In the<br />

first generation, one mother, Nippy, and her calf, Nassau,<br />

demonstrated the presumed teaching behavior. In the second<br />

generation, Nassau, now a mother, showed the same<br />

altered foraging behavior as her mother with her calf,<br />

Neptune.<br />

Foraging with non-calf individuals<br />

Of the foraging events without calves, three mothers were<br />

observed foraging both alone and with other individuals<br />

that were not calves, at least juveniles or older. The chase<br />

latencies and body-orienting movements for these mothers<br />

were compared for the two conditions (alone and with noncalf<br />

individuals). There were too few subjects (n = 3) to<br />

perform a statistical test; however, the chase latencies of<br />

these three mothers foraging alone (M = 1.33, SD = 0.58)<br />

were comparable to the chase latencies of these same<br />

mothers foraging with non-calf individuals (M = 3.44,<br />

SD = 1.50), and both were much lower than the chase<br />

24.33<br />

2.25 2.00 2.00 2.<strong>17</strong><br />

Mother<br />

19.67<br />

4.00<br />

Mothers Chasing<br />

With Calves<br />

Mothers Chasing<br />

Without Calves<br />

Fig. 3 Mean chase latencies of mothers foraging with and without<br />

their calves present<br />

16.50<br />

1.67<br />

12.00<br />

4.00<br />

14.00<br />

1.00<br />

123


48 Anim Cogn (2009) 12:43<strong>–</strong>53<br />

latencies of those mothers foraging with their calves<br />

present (M = <strong>17</strong>.56, SD = 6.26). Additionally, the number<br />

of body-orienting movements of these three mothers foraging<br />

alone (M = 0.<strong>17</strong>, SD = 0.29) was comparable to the<br />

number of body-orienting movements when they were<br />

foraging with non-calf individuals (M = 0.33, SD = 0.58),<br />

and both were less than the number of body-orienting<br />

movements of those mothers foraging with their calves<br />

present (M = 1.94, SD = 0.42).<br />

Additional analyses<br />

The differences observed for chase latencies and number of<br />

body-orienting movements could not be attributed to prey<br />

type, as there were no significant differences in the foraging<br />

behaviors between prey species. Prey species were identified<br />

for 14 observations. When foraging with calves present,<br />

mothers were observed foraging for Snakefish (family<br />

Synodontidae), n = 5, flounder (family Bothidae), n = 2,<br />

and razorfish (family Clinidae), n = 1. The mean chase<br />

latencies, M Snakefish = 19.0 (SD = 11.03), M Flounder = 20.5<br />

(SD = 14.85), MRazorfish = 25.0 (SD = 0.0), and number<br />

of body-orienting movements, MSnakefish = 0.6 (SD =<br />

0.89), M Flounder = 25.0 (SD = 2.12), M Razorfish = 3.0<br />

(SD = 0.0), were comparable for all three species of prey<br />

when foraging with calves present. Both snakefish, n = 9,<br />

and flounder, n = 4, were observed as prey types when the<br />

females were observed foraging without calves present. The<br />

mean chase latencies, MSnakefish = 2.22 (SD = 1.09),<br />

M Flounder = 3.83 (SD = 3.57), and number of body-orienting<br />

movements, MSnakefish = 0.5 (SD = 0.76), MFlounder =<br />

0.5 (SD = 0.58), were comparable for both species when<br />

foraging without calves present, and both were lower than<br />

when foraging with calves present.<br />

Additionally, individual foraging events of dolphins not<br />

included in the study observed foraging for either snakefish<br />

or flounder were collected and analyzed. Ten foraging<br />

events each for catches of snakefish and flounder using<br />

both dolphins in the present study and dolphins not used in<br />

the study were randomly selected, and their corresponding<br />

chase latencies and number of body-orienting movements<br />

were compared between the two types of prey using an<br />

independent samples Student’s t-test. There were no significant<br />

differences found between the two types of prey for<br />

either chase latencies, (M flounder = 4.60, SD flounder = 3.86,<br />

Msnakefish = 2.80, SDsnakefish = 1.62) t18 =-1.36,<br />

P = 0.19, or number of body orienting movements for<br />

dolphins foraging without calves present, (M flounder = 0.50,<br />

SDflounder = 0.71, Msnakefish = 0.37, SDsnakefish = 0.48)<br />

t18 =-0.49, P = 0.63. This supports the conclusion that<br />

the observed differences in this study for mothers foraging<br />

with calves present would not likely be due to the type of<br />

prey, as chase latencies and the number of body-orienting<br />

123<br />

movements do not normally vary significantly between<br />

snakefish and flounder, the two main types of prey observed<br />

being caught by mother dolphins in this study.<br />

To assess possible effects of age of calf and age of<br />

mother on the dependent measures, correlations with each<br />

dependent measure were computed separately with the age<br />

of the calf and the age of the mother for the foraging events<br />

when the calf was present. Correlations between age of the<br />

calf and chase latency, r13 = 0.37, P = 0.10, and number<br />

of body-orienting movements, r13 =-0.30, P = 0.15,<br />

were both nonsignificant. Correlations between mother’s<br />

age and number of body-orienting movements were significant,<br />

r13 =-0.47, P = 0.046, with older mothers<br />

making fewer body-orienting movements than younger<br />

mothers. The correlation between mother’s age and chase<br />

latencies was not significant, r13 = 0.01, P = 0.48. When<br />

comparing the foraging behaviors to the age of the mothers<br />

when the calves were not present, correlations with both<br />

chase latencies r23 =-0.07, P = 0.73, and number of<br />

body-orienting movements, r 23 =-0.12, P = 0.58, were<br />

nonsignificant. The foraging behaviors of eight of the nine<br />

mothers without calves were compared between those<br />

females observed as juveniles and those observed as adults,<br />

with three mothers being observed foraging as juveniles<br />

and five mothers observed foraging as adults. 1 There were<br />

no significant differences between chase latencies, t (6) =-<br />

0.10, P = 0.92, or number of body-orienting movements,<br />

t(6) =-0.12, P = 0.91, of females observed foraging<br />

when juveniles or adults.<br />

Additionally, of the nine mothers observed in this study,<br />

three of the mothers, Little Gash, Mugsy, and PR2, were<br />

observed foraging both with and without their calves<br />

present during the same year, and each of those three<br />

mothers being observed foraging both with and without<br />

calves during the same field encounter. The mean chase<br />

latencies for these three mothers are compared in Table 2.<br />

The chase latencies were much longer foraging with calves<br />

than foraging alone for the mothers at the same age and<br />

comparable to the mean chase latencies of foraging alone<br />

for the mothers at an earlier age.<br />

The position of the calf during the chase period was<br />

indicated for each of the foraging events (N = 14) in which<br />

the calf was present. Although some foraging events<br />

involved multiple calf positions, the observation position<br />

1 One mother, Little Gash, was observed foraging without her calf<br />

present both as an adult and as a juvenile. When group means were<br />

compared including Little Gash there were no significant differences<br />

between adults and juveniles for either chase latencies, t(8) = 0.05,<br />

P = 0.96, or body-orienting movements, t (8) = 0.41, P = 0.69. Both<br />

the mean chase latencies, M A = 3.67 and M J = 6.89, and mean<br />

number of body-orienting movements, M A = 0.5 and M J = 1.0, for<br />

Little Gash were comparable as an adult and as a juvenile.


Anim Cogn (2009) 12:43<strong>–</strong>53 49<br />

Table 2 Maternal age<br />

comparison for mean chase<br />

latency and number of bodyorienting<br />

movements<br />

was found to be the most common, taking place in 11<br />

(79%) of the foraging events. Head-under-head position<br />

was present in one of the foraging events, Echelon was<br />

present in two foraging events, and Infant position was<br />

seen in one foraging event. Only Infant, Echelon, and<br />

Observation positions were seen solely in an individual<br />

foraging event, with one, one, and nine occurrences,<br />

respectively.<br />

For the foraging events collected in which the ingestion<br />

by the mother was known (as seen by video or first-hand<br />

account), there were five events, one event each for five of<br />

the nine mothers, in which the prey was not eaten. Each of<br />

these foraging events occurred when the mothers’ calves<br />

were present (36% of the total events with calves). In<br />

addition, the calves were allowed to pursue the prey in each<br />

of these events and were confirmed to have eaten the prey<br />

in three of the foraging events, despite the fact that they<br />

were still nursing and not dependent upon fish for food.<br />

Four of the nine mothers were observed eating the prey in<br />

seven of the fourteen events in which calves were present.<br />

In two of the events in which calves were present it was not<br />

known whether the prey was eaten.<br />

Discussion<br />

Encounter dates<br />

by mother<br />

The nine mother dolphins observed in this study displayed<br />

significantly longer chase latencies and made significantly<br />

more body-orienting movements when foraging in the<br />

presence of their calves than when foraging alone. The<br />

chase latency data are particularly impressive. Mean chase<br />

latencies were eight times longer for the female dolphins<br />

when foraging with their calves (22.24 s) than without<br />

them (2.74 s). As illustrated in Fig. 3, the distributions<br />

were non-overlapping, with every mother having a longer<br />

latency when foraging with her calf than when foraging<br />

without her calf.<br />

Although differences were not as robust for the bodyorienting<br />

movements, the overall number of body-orienting<br />

Mean chase latency<br />

with calf (n)<br />

Mean chase latency<br />

without calf (n)<br />

Mean BOM<br />

with calf<br />

Mean BOM<br />

without calf<br />

Little Gash<br />

30 May 2000 38.33 (1) 3.67 (2) 0.00 (1) 0.50 (2)<br />

27 May 1991<br />

Mugsy<br />

<strong>–</strong> 4.00 (2) <strong>–</strong> 1.50 (2)<br />

29 August 1993 34.00 (1) 2.00 (1) 5.00 (1) 0.00<br />

13 May 2000 <strong>–</strong> 4.00 (1) <strong>–</strong> 0.00<br />

11 August 2000<br />

PR2<br />

12.00 (1) <strong>–</strong> 0.00 <strong>–</strong><br />

14 June 1994 19.67 (1) 4.00 (1) 2.00 0.00<br />

movements was significantly greater when the mothers<br />

were foraging with their calves than when foraging alone,<br />

with six of nine mothers displaying this pattern. Previous<br />

research has shown that dolphins are capable of understanding<br />

the human gesture of pointing (Herman et al.<br />

1999; Pack and Herman 2004), and of producing spontaneous<br />

referential gestures in artificial experimental<br />

contexts (Xitco et al. 2001). Although these body-orienting<br />

movements were not demonstrated by all sampled mothers,<br />

the present results may be evidence of referential gesturing<br />

in a more ecologically valid context.<br />

In addition to the elongated chase and presumed referential<br />

body-orienting movements in the direction of the<br />

prey, the altered foraging behavior in the presence of the<br />

calf appeared more exaggerated when compared to the<br />

mothers foraging alone or with another non-calf dolphin.<br />

Mothers seemed to toy with their prey, making it more like<br />

play behavior and less like the typical foraging behavior of<br />

mothers observed foraging without their calves present,<br />

similar to descriptions of possible teaching of predatory<br />

behavior from other species such as cats and killer whales<br />

(Caro and Hauser 1992; Guinet and Bouvier 1995). Some<br />

of the mothers were observed letting the prey swim away<br />

and then digging them back out of the sand, sometimes<br />

multiple times, after initially digging the fish out of the<br />

sand, and also make jawing motions in the direction of the<br />

prey. Mothers also allowed the seemingly attentive calves<br />

to participate in the chase, and calves were observed eating<br />

the prey in three of the events. Although 4 of the 9 mothers<br />

were observed eating the prey in 7 of the 14 events, all of<br />

the mothers were observed allowing the calves to participate<br />

and made little to no effort in these events to consume<br />

the prey, only seemingly facilitating the calves’ experience<br />

in chasing the prey. Despite the altered foraging behavior,<br />

mothers were never observed losing the prey. In all events<br />

in which ingestion of the prey was observed, either the<br />

mother or the calf ate the prey, or the mother made no<br />

further attempts to ingest the prey and left the prey to the<br />

calf to chase. This altered foraging behavior can be<br />

123


50 Anim Cogn (2009) 12:43<strong>–</strong>53<br />

observed in the online supplemental movies S1 and S2<br />

which exemplify the markedly different foraging behavior<br />

of an adult female foraging alone (S1) with the foraging<br />

behavior of a mother foraging in the presence of her calf<br />

(S2).<br />

There was a potential cost of the alteration of the<br />

mothers’ foraging behavior in their calves’ presence. The<br />

elongated chase time allowed more opportunity for prey to<br />

escape, although this did not occur in any of the events<br />

observed, as well as taking time away from catching<br />

additional fish or other activities. Instead of the typical grab<br />

and ingestion, these mothers toyed with their prey as their<br />

calves watched, and, as mentioned above, sometimes even<br />

avoided ingestion in these altered foraging events. Often in<br />

teaching, the model will provide an exaggerated or elongated<br />

version of the typical behavior in front of the naïve<br />

observer, much as the mother chimpanzees observed in<br />

Boesch’s (1991) work. These exaggerated foraging<br />

behaviors may provide a window of opportunity for the<br />

calves to observe, and possibly learn from, the example<br />

provided by their mothers, and thus be worthy of the extra<br />

time and energy put forth by the mothers at a cost to<br />

themselves in order to help ensure the success of their<br />

offspring.<br />

An alternative explanation for the altered foraging<br />

behavior may be that the mothers were distracted by their<br />

calves, resulting in longer chase latencies and exaggerated<br />

movements to compensate for their divided attention.<br />

However, we do not believe this to be the cause of the<br />

altered foraging behavior for a few reasons. First, three of<br />

the nine mothers observed in the present study demonstrated<br />

the ability of dolphin mothers to forage without<br />

distraction even when their calves were in the nearby<br />

vicinity, but not directly observing the mothers. These<br />

three mothers, Little Gash, Mugsy, and PR2, were<br />

observed foraging both with and without their calves<br />

present during the same field encounter. These mothers<br />

only altered their behavior when the calves were<br />

observing the mother, but not during the fish catches in<br />

which the calves were within the vicinity and not<br />

observing. Table 2 compares the means for those events,<br />

in which the mean chase latencies were much longer<br />

foraging with calves than foraging alone during the same<br />

field encounter and comparable to the mean chase latencies<br />

of foraging alone during separate field encounters.<br />

These data suggest that the calves’ being merely nearby is<br />

not a distraction for the mothers, but instead the altered<br />

foraging behavior only occurred when the calves were<br />

directly observing the mothers’ behavior. If the calves<br />

were a distraction for the mothers, the altered foraging<br />

behavior should have occurred whenever the calves were<br />

within the vicinity, regardless of whether the calves were<br />

directly observing the mothers.<br />

123<br />

Second, Atlantic spotted dolphins participate in alloparenting,<br />

which permits mothers to frequently forage<br />

separately from their calves (Herzing 1996). If the mother’s<br />

nutritional needs are being met in the absence of their calf<br />

with the help of alloparenting, this would enable more time<br />

and energy on the part of the mother for altered foraging<br />

events for teaching when the calf is present. Additionally, it<br />

would not be advantageous for the mother to forage with<br />

the calf present solely for her nutritional purposes if there is<br />

an alloparent available. In this study, there were potential<br />

alloparents, juvenile or adult female dolphins, available in<br />

all 14 of the calf-present events that would have permitted<br />

the mothers to forage without distraction, which is often<br />

observed in Atlantic spotted dolphins (Herzing 1996),<br />

including the events for the previously mentioned three<br />

mothers observed foraging with and without calves during<br />

the same field encounter.<br />

Third, young dolphin calves are very precocious and the<br />

calves in the observed foraging events appeared attentive<br />

and interactive. Some of the evidence for the calves’<br />

attentiveness to the mothers’ foraging behavior was<br />

observed through the calves’ positions relative to the<br />

mothers, which was indicated during the chase period for<br />

each of the foraging events in which the calf was present.<br />

Although previous research (Mann and Smuts 1999) has<br />

shown that dolphin calves in the wild, and in captivity<br />

spend the majority of their time in infant or echelon position<br />

until the time of weaning, calves were most commonly<br />

found during the chase period in the observation position<br />

relative to their mothers in which the calf was potentially<br />

exposed to both visual and acoustic information and in<br />

which calves appeared attentive to both the mother and the<br />

prey 2 (Figure 2).<br />

Additionally, the calves were allowed to pursue the prey<br />

in five of the foraging events and were confirmed to have<br />

eaten the prey in three of the events, despite the fact that<br />

they were still nursing and not dependent upon fish for<br />

food. It is important to note that the only time when the fish<br />

was not ingested was when the mothers were foraging with<br />

their calves as part of their altered foraging behavior. The<br />

prey was always consumed when foraging alone or in the<br />

presence of a non-calf dolphin. This is potentially a costly<br />

behavior for the mother as she is depriving herself of calories<br />

needed to nourish her still nursing calf by allowing<br />

2 It is important to note that research in captivity (Pack and Herman<br />

1995) has demonstrated the ability of dolphins to perceive and<br />

recognize objects through either vision or echolocation. In addition,<br />

their perceptions are readily shared or integrated across the senses,<br />

regardless of which modality the dolphin originally perceived the<br />

external stimuli. However, the sounds emitted from the dolphins were<br />

not measured in the present study, but should be looked at in future<br />

research in order to determine what sensory information the calf is<br />

receiving.


Anim Cogn (2009) 12:43<strong>–</strong>53 51<br />

the prey to escape or allowing the calf to eat the prey, as<br />

well as the energy to play with the prey rather than just<br />

eating it quickly and efficiently herself. It would be more<br />

efficient and less costly for the mothers to have simply<br />

caught and consumed the fish, or perhaps to have left the<br />

calf in the care of an alloparent, as opposed to this altered<br />

foraging behavior observed in these events.<br />

Furthermore, the chase latencies and foraging behaviors<br />

were drastically different with their calves present, not just<br />

with the longer latencies but also with the presumed referential<br />

behaviors toward the prey objects and allowing the<br />

still nursing calf to participate, which would be more<br />

indicative of social learning. If the calves were merely a<br />

distraction, the mothers could have either immediately<br />

consumed the prey or disciplined the calf, as previously<br />

observed by mothers in the population for undesirable<br />

behaviors (Miles and Herzing 2003), rather than allowing<br />

the calf to participate. Therefore, because the calves<br />

appeared attentive, interacted with the prey and the mother,<br />

and appeared to need little care during the event, they were<br />

not likely a distraction to the mother.<br />

Reciprocally, the mothers may have been altering their<br />

foraging behavior because their calves were attentive, so<br />

that a calf’s attention may have stimulated the altered<br />

maternal foraging behavior. Further analysis into the<br />

calves’ behavior is warranted and may also elucidate the<br />

exact learning mechanism at play on behalf of the calf.<br />

For the majority of the mothers, there was a difference<br />

in maternal age between the events observed foraging<br />

without their calves compared to the events foraging with<br />

their calves. It is not likely that the age differences, or<br />

resulting level of experience would have resulted in the<br />

observed differences in chase latencies and number of<br />

body-orienting movements. Rather, an older or more<br />

experienced dolphin should be expected to have quicker<br />

chase latencies and fewer body-orienting movements due<br />

to increased efficiency, not the longer latencies and<br />

increased number of body-orienting movements as<br />

observed here. Additionally, of the nine mothers observed<br />

in this study, three of the mothers, Little Gash, Mugsy, and<br />

PR2, were observed foraging both with and without their<br />

calves present during the same year, and each of these<br />

three mothers were observed foraging both with and<br />

without calves during the same field encounter. For these<br />

three mothers, mean chase latencies were much longer<br />

foraging with calves than foraging alone at the same age<br />

and comparable to the mean chase latencies of foraging<br />

alone at an earlier age (Table 2). This suggests that there<br />

was not a difference in foraging behavior due to difference<br />

in age or experience, but rather due to the presence of their<br />

calves.<br />

There were also no significant differences for chase<br />

latencies or number of body-orienting movements for the<br />

prey species observed in this study: thus the altered foraging<br />

behaviors were also not likely due to the difficulty of<br />

catching a specific type of prey.<br />

The three episodes in which a target dolphin was<br />

observed foraging with a non-calf individual, juvenile in<br />

age or older, produced chase latencies and number of bodyorienting<br />

movements comparable to when these dolphins<br />

foraged alone, and both substantially different from the<br />

levels of behavior observed when these dolphins foraged<br />

with their calves present. This suggests that the change in<br />

behavior was not merely a social phenomenon seen in the<br />

presence of other individuals, but instead reflects teaching<br />

behavior targeted at a naïve observer. However, the naïve<br />

observers in this study were presumably the calves of the<br />

observed mothers. Future research is needed to explore if<br />

these same mothers or other alloparents also exhibit similar<br />

altered foraging behavior when in the presence of other<br />

naïve observers that are not their offspring.<br />

This altered foraging behavior on behalf of the mothers<br />

may be a valuable social learning mechanism for Atlantic<br />

spotted dolphins. The mothers in the study clearly altered<br />

their typical foraging behavior in the presence of their<br />

calves at a potential cost to themselves due to the exaggerated<br />

behavior, as well as sometimes foregoing ingestion<br />

of the prey. However, as per Caro and Hauser’s definition<br />

of teaching, the observer must benefit from the model’s<br />

altered behavior, in this case through more rapid or skillful<br />

acquisition of foraging behavior. In dolphin society this<br />

skill would be essential, as mothers give birth approximately<br />

every 3 years, at which time the older calf would<br />

become weaned and rely more on independently caught<br />

fish. It is advantageous for the mother to ensure the success<br />

of her offspring, presumably by investing her time and<br />

energy into the calf’s foraging capabilities before the calf<br />

is weaned. Weaning is a gradual process in spotted dolphins,<br />

and although the calves were not dependent upon<br />

fish for food, consuming the fish may have been an<br />

important part of the development of the calves’ foraging<br />

behavior, perhaps because it reinforced the social-learning<br />

process.<br />

Future research is warranted to explore the development<br />

and skillfulness of the young calf’s pre-weaning foraging<br />

capabilities in order to examine the full effect of the<br />

mother’s presumed teaching efforts. If this is true teaching<br />

behavior, the calf will derive some benefit from observing<br />

the mother’s altered foraging behavior. Additionally, a<br />

comparison is needed to compare the calves of teaching<br />

and non-teaching or less attentive mothers. However,<br />

because all of the mothers observed in this study demonstrated<br />

the altered foraging behavior, it may be difficult to<br />

assess the consequences of naturally occurring individual<br />

differences in the mothers’ foraging behavior on the calves.<br />

Although the full benefit for the calf is unknown, it seems<br />

123


52 Anim Cogn (2009) 12:43<strong>–</strong>53<br />

that the calf is indeed the target of this altered behavior,<br />

which was observed only in the presence of the calf.<br />

Additionally, further research is needed into the calf’s<br />

behavior and attention to the mother. Future research can<br />

hopefully clarify if the calf is indeed attentive to the<br />

mother’s potential teaching behavior to gain something<br />

from the experience, as well as clarify what mechanism<br />

the calves may be using to learn from the mothers, such<br />

as imitation, stimulus enhancement, or local enhancement.<br />

Data from this study, such as the calves being<br />

predominantly in the observation position relative to the<br />

mother during the events, some of the calves chasing and<br />

ingesting the fish, and the calves’ interactions with both<br />

the mothers and prey objects, may be evidence that the<br />

calves were attentive to the mothers’ altered foraging<br />

behavior and support the argument for teaching. Future<br />

data from the calf behavior will also hopefully<br />

strengthen our argument that the altered maternal foraging<br />

behavior may be an example of teaching. The data<br />

from the calf behavior are currently being collected and<br />

analyzed as part of a separate ongoing project and will<br />

be available for future publication. Despite this drawback,<br />

we believe that this study detailing the altered<br />

foraging behavior of the mothers is a significant finding<br />

in the area of animal cognition, even without data on the<br />

calf behavior.<br />

Despite previous research that has shown that dolphins<br />

pass mirror self-recognition tests (Reiss and Marino 2001),<br />

understand referential pointing (Herman et al. 1999), and<br />

spontaneously use referential gesturing in captive situations<br />

(Xitco et al. 2001), further evidence is needed before<br />

attributing theory of mind to dolphins, which some<br />

researchers argue is required for true teaching (Tomasello<br />

et al. 1993). Regardless of the lack of conclusive evidence<br />

supporting theory of mind in dolphins, the perspectivetaking<br />

abilities by dolphins supported by previous research<br />

(Herman et al. 1999; Xitco et al. 2001) might be sufficient<br />

for the presumed teaching behavior shown here. Although<br />

the cognitive abilities behind the clear alteration of foraging<br />

behavior of the mother dolphins in the presence of their<br />

calves are yet to be determined, the observed teaching<br />

behaviors in dolphins are nonetheless remarkable. Mother<br />

dolphins provide opportunities for calves to observe foraging<br />

behaviors, and sometimes even provide opportunities<br />

for calves to practice foraging. Teaching, then, may be an<br />

important way in which aspects of cetacean social learning<br />

and possibly culture are transmitted from one generation to<br />

the next.<br />

Acknowledgments We would like to thank Stan Kuczaj and Jesse<br />

Bering for their helpful comments on earlier drafts of this manuscript,<br />

and John Bender, Miley Fishero, Megan Rothrock, Melissa Ingui,<br />

Wisline Shepherd, Sheryl Spencer, and the Wild Dolphin Project for<br />

their assistance on the project.<br />

123<br />

References<br />

Bering JM (2001) Theistic percepts in other species: can chimpanzees<br />

represent the minds of non-natural agents? J Cogn Cult 1:107<strong>–</strong><br />

137<br />

Bering JM, Povinelli DJ (2003) Comparing cognitive development.<br />

In: Maestripieri D (ed) Primate psychology. Harvard University<br />

Press, Cambridge, pp 205<strong>–</strong>233<br />

Bjorklund DF, Bering JM (2003) Big brains, slow development, and<br />

social complexity: the developmental and evolutionary origins of<br />

social cognition. In: Brüne M, Ribbert H, Schiefenhövel W (eds)<br />

The social brain: evolutionary aspects of development and<br />

pathology. Wiley, New York, pp 133<strong>–</strong>151<br />

Bjorklund DF, Yunger JL, Bering JM, Ragan P (2002) The<br />

generalization of deferred imitation in enculturated chimpanzees<br />

(Pan troglodytes). Anim Cogn 5:49<strong>–</strong>58<br />

Boesch C (1991) Teaching in wild chimpanzees. Anim Behav<br />

41:530<strong>–</strong>532<br />

Boesch C (1993) Toward a new image of culture in chimpanzees.<br />

Behav Brain Sci 16:514<strong>–</strong>515<br />

Boesch C, Tomasello M (1998) Chimpanzee and human cultures.<br />

Curr Anthropol 39:591<strong>–</strong>614<br />

Caro TM, Hauser MD (1992) Is there teaching in nonhuman animals?<br />

Q Rev Biol 67:151<strong>–</strong><strong>17</strong>4<br />

Greenfield P, Maynard A, Boehm C, Schmidtling EY (2000) Cultural<br />

apprenticeship and cultural change: tool learning and imitation in<br />

chimpanzees and humans. In: Parker ST, Langer J, McKinney<br />

ML (eds) Biology, brains, and behavior: the evolution of human<br />

development. School of American Research Press, Santa Fe, pp<br />

237<strong>–</strong>277<br />

Guinet C, Bouvier J (1995) Development of intentional stranding<br />

hunting techniques in killer whale (Orcinus orca) calves at<br />

Crozet Archipelago. Canadian Journal of Zoology 73:27<strong>–</strong>33<br />

Herman LM (2002) Vocal, social, and self-imitation by bottlenosed<br />

dolphins. In: Nehaniv C, Dautenhahn K (eds) Imitation in<br />

animals and artifacts. MIT Press, Cambridge, pp 63<strong>–</strong>108<br />

Herman LM, Abichandani SL, Elhajj AN, Herman EYK, Sanchez JL,<br />

Pack AA (1999) Dolphins (Tursiops truncatus) comprehend the<br />

referential character of the human pointing gesture. J Comp<br />

Psychol 113:347<strong>–</strong>364<br />

Herzing DL (1996) Vocalizations and associated underwater behavior<br />

of free-ranging Atlantic spotted dolphins, Stenella frontalis, and<br />

bottlenose dolphins, Tursiops truncatus. Aquat Mamm 22:61<strong>–</strong>79<br />

Herzing DL (1997) The life history of free-ranging Atlantic spotted<br />

dolphins (Stenella frontalis): age classes, color phases an female<br />

reproduction. Mar Mamm Sci 13:576<strong>–</strong>595<br />

Herzing DL (2005) Transmission mechanisms of social learning in<br />

dolphins: underwater observations of free-ranging dolphins in<br />

the Bahamas. In: Delfour F (ed) Autour de L’ethologie et de la<br />

cognition animale. Presses Universitaires de Lyon, Lyon, pp<br />

185<strong>–</strong>194<br />

Herzing DL, Brunnick BJ (1997) Coefficients of association of<br />

reproductively active female Atlantic spotted dolphins, (Stenella<br />

frontalis). Aquat Mamm 23:155<strong>–</strong>162<br />

Kruetzen M, Mann J, Heithaus M, Connor R, Bejder L, Sherwin WB<br />

(2005) Cultural transmission of tool use in bottlenose dolphins.<br />

Proc Natl Acad Sci USA 105:8939<strong>–</strong>8943<br />

Kuczaj SAII, Paulos RD, Ramos JA (2005) Imitation in apes, children<br />

and dolphins: implications for the ontogeny and phylogeny of<br />

symbolic representation. In: Namy L (ed) Symbol use and<br />

symbolic development. MIT Press, Cambridge, pp 221<strong>–</strong>243<br />

Mann J, Smuts BB (1999) Behavioral development in wild bottlenose<br />

dolphin newborns (Tursiops sp.). Behaviour 136:529<strong>–</strong>566<br />

Marino L, Connor RC, Fordyce RE, Herman LM, Hof PR, Lefebvre<br />

L, Lusseau D, McCowan B, Nimchinsky EA, Pack AA, Rendell


Anim Cogn (2009) 12:43<strong>–</strong>53 53<br />

L, Reidenberg JS, Reiss D, Uhen MD, Van der Gucht E,<br />

Whitehead H (2007) Cetaceans have complex brains for<br />

complex cognition. PLoS Biol 5:966<strong>–</strong>972<br />

Miles JA, Herzing DL (2003) Underwater analysis of the behavioural<br />

development of free-ranging Atlantic spotted dolphin (Stenella<br />

frontalis) calves (birth to 4 years of age). Aquat Mamm 29:363<strong>–</strong><br />

377<br />

Pack AA, Herman LM (1995) Sensory integration in the bottlenosed<br />

dolphin: immediate recognition of complex shapes across the<br />

senses of echolocation and vision. J Acoust Soc Am 98:722<strong>–</strong>733<br />

Pack AA, Herman LM (2004) Dolphins (Tursiops truncatus) comprehend<br />

the referent of both static and dynamic human gazing and<br />

pointing in an object choice task. J Comp Psychol 118:160<strong>–</strong><strong>17</strong>1<br />

Reiss D, Marino L (2001) Self-recognition in the bottlenose dolphin:<br />

a case of cognitive convergence. Proc Natl Acad Sci USA<br />

98:5937<strong>–</strong>5942<br />

Rendell L, Whitehead H (2001) Culture in whales and dolphins.<br />

Behav Brain Sci 24:309<strong>–</strong>382<br />

Spininelli LHP, Jesus AH, Nascimento LF, Yamamoto ME (2006)<br />

Prey-transfer in the marine tucuxi dolphin, Sotalia fluviatilis, on<br />

the Brazilian coast. JMBA2, Biodiversity Records (published<br />

online)<br />

Thornton A, McAuliffe K (2006) Teaching in wild meerkats. Science<br />

313:227<strong>–</strong>229<br />

Tomasello M (1996) Do apes ape? In: Heyes C, Galef B (eds) Social<br />

learning in animals: the role of culture. Academic Press, San<br />

Diego, pp 319<strong>–</strong>346<br />

Tomasello M (2000) Culture and cognitive development. Curr Dir<br />

Psychol Sci 9:37<strong>–</strong>40<br />

Tomasello M, Kruger AC, Ratner HH (1993) Cultural learning.<br />

Behav Brain Sci 16:495<strong>–</strong>552<br />

Whiten A (2005) The second inheritance system of chimpanzees and<br />

humans. Nature 437:52<strong>–</strong>55<br />

Whiten A, Goodall J, McGrew WC, Nishida T, Reynolds V,<br />

Sugiyama Y, Tutin CEG, Wrangham RW, Boesch C (1999)<br />

Cultures in chimpanzees. Nature 399:682<strong>–</strong>685<br />

Xitco MJ, Gory JD, Kuczaj SAII (2001) Spontaneous pointing by<br />

bottlenose dolphins (Tursiops truncatus). Anim Cogn 4:115<strong>–</strong>123<br />

123


Aquatic Mammals 2009, 35(1), 62-71, DOI 10.1578/AM.35.1.2009.62<br />

Characterization of Resting Holes and Their Use by the<br />

Antillean Manatee (Trichechus manatus manatus)<br />

in the Drowned Cayes, Belize<br />

Marie-Lys C. Bacchus, 1, 2 Stephen G. Dunbar, 1, 2 3, 4<br />

and Caryn Self-Sullivan<br />

1 Department of Earth and Biological Sciences, Loma Linda University, Loma Linda, CA 92350, USA;<br />

E-mail: mlbacchus@gmail.com<br />

2 Marine Research Group, Loma Linda University, Loma Linda, CA 92350, USA<br />

3 Texas A&M University, Department of Wildlife and Fisheries, College Station, TX 77843, USA<br />

4 <strong>Sirenian</strong> <strong>International</strong>, Inc., 200 Stonewall Drive, Fredericksburg, VA 22401, USA<br />

Abstract<br />

In the Drowned Cayes area of Belize, manatees<br />

(Trichechus manatus manatus) are commonly<br />

observed resting in depressions in the substrate,<br />

locally referred to as manatee resting holes. To<br />

understand why manatees prefer locations with<br />

resting holes, the physical and environmental<br />

attributes of the depressions were characterized<br />

and diurnal and nocturnal use by manatees at four<br />

resting hole sites were documented over two summers.<br />

Twelve resting hole sites were compared<br />

with 20 non-resting hole sites in the Drowned<br />

Cayes, using water depth, substrate type, vegetation,<br />

water velocity, salinity, and water temperature.<br />

Four resting holes were chosen for repeated<br />

diurnal and nocturnal observations, during which<br />

sea and weather conditions were recorded in addition<br />

to the presence/absence of manatees. Resting<br />

holes were significantly deeper and had slower<br />

surface water velocity than areas without resting<br />

holes. A total of 168 point scans were conducted<br />

over 55 d, resulting in 39 manatee sightings over<br />

two summers. There was a significant difference<br />

in the number of sightings between research years<br />

and between day and night scans. Given the large<br />

number of resting holes in the Drowned Cayes,<br />

many of which are in sheltered areas with slow<br />

currents, it is possible that manatees select these<br />

spots based on the tranquility of the water and<br />

environment. The combination of slow currents,<br />

protection from waves, low numbers of boats, and<br />

nearby seagrass beds would make these ideal resting<br />

areas. These findings have implications for the<br />

conservation of important manatee habitat.<br />

Key Words: Antillean manatee, Trichechus<br />

manatus manatus, Belize, resting holes, nocturnal,<br />

diurnal, conservation, habitat, sightings<br />

Introduction<br />

The West Indian manatee (Trichechus manatus)<br />

is one of only four extant species in the Order<br />

Sirenia, which is the only group of herbivorous<br />

marine mammals. The species is subdivided into<br />

Florida (T. m. latirostris) and Antillean (T. m.<br />

manatus) manatees (Domning & Hayek, 1986).<br />

The 2007 IUCN Red List categorized the species<br />

as “vulnerable” and both subspecies as “endangered”<br />

(Deutsch et al., 2007).<br />

West Indian manatees inhabit rivers, lakes,<br />

lagoons, and coastal marine environments from<br />

southeastern North America to northeastern<br />

South America, including Central America and<br />

the Caribbean (Lefebvre et al., 2001; Deutsch<br />

et al., 2007). Within Central America, aerial<br />

surveys have indicated that the cayes east of<br />

Belize City are important manatee habitat (Auil,<br />

2004). Manatees osmoregulate and thermoregulate<br />

behaviorally by moving between activity<br />

centers (Deutsch et al., 2007). They can survive<br />

in marine environments for extended periods of<br />

time; however, a dependence on periodic access to<br />

fresh water is presumed (Ortiz et al., 1998, 1999;<br />

Deutsch et al., 2003). Thus, when found in marine<br />

or coastal environments, they tend to be located<br />

near freshwater sources (Powell et al., 1981;<br />

Powell & Rathbun, 1984; Rathbun et al., 1990;<br />

Olivera-Gomez & Mellink, 2005). Proximity to<br />

seagrass beds is another important characteristic in<br />

determining manatee habitat preference (Hartman,<br />

1979; Powell et al., 1981; Deutsch et al., 2003). In<br />

Chetumal Bay, Mexico, manatee movements were<br />

most strongly associated with food distribution<br />

(Axis-Arroyo et al., 1998), and in Florida, aerial<br />

surveys indicated a strong association between<br />

location of manatees and distribution of seagrass<br />

beds. Work by Kinnaird (1985) and Provancha<br />

& Hall (1991), showed manatee density was<br />

positively correlated with seagrass abundance


(Halodule wrightii and Syringodium filiforme) in<br />

the northern Banana River, Florida. Some of the<br />

most important manatee food items around the<br />

cayes of Belize are turtle grass (Thalassia testudinum),<br />

manatee grass (S. filiforme), and shoal grass<br />

(H. wrightii) (LaCommare et al., 2008). There<br />

is also evidence that resting areas for cows and<br />

calves, as well as travel corridors between activity<br />

centers, may be critical habitats for manatees<br />

(Packard & Wetterqvist, 1986; Deutsch et al.,<br />

2007). Major threats to survival of the species<br />

include habitat degradation and loss, illegal hunting,<br />

boat strikes, entanglement in fishing gear,<br />

entrapment in water control structures, pollution,<br />

disease, and human disturbance (Deutsch et al.,<br />

2007).<br />

Conservation of this species is dependent on<br />

conservation of habitat that meets the requirements<br />

described above. Compared to osmoregulation,<br />

thermoregulation, and foraging requirements,<br />

little is known about manatee resting<br />

behavior and the importance of resting habitat<br />

for conservation of the species. Hartman (1979)<br />

and Bengtson (1981) described Florida manatees<br />

as arrhythmic, spending most of their time feeding,<br />

resting, idling, cruising, and socializing with<br />

no correlation of activity with time of day. The<br />

hypothesis that Florida manatees lack a daily<br />

activity pattern is supported by two lines of evidence:<br />

(1) similar nocturnal and diurnal behaviors<br />

(Hartman, 1971) and (2) lack of a pineal body<br />

(Ralph et al., 1985; Wally Welker, pers. comm.).<br />

Hartman (1979) also found that Florida manatees<br />

devoted considerable time to resting behavior<br />

with no strict preference for resting sites, which<br />

included limestone shelves, oyster bars, and vegetation.<br />

However, more recent studies indicated<br />

that manatees have a diurnal or nocturnal preference<br />

for resting in different areas of Florida. For<br />

example, during winter along the Atlantic coast,<br />

manatees rested in canals during the day and foraged<br />

at night (Deutsch et al., 2003). In Sarasota,<br />

they will bottom-rest in deeper areas during the<br />

winter when it is colder, and then will switch to<br />

surface resting when temperatures increase (Sheri<br />

W. Barton, pers. comm.). Along the Gulf Coast<br />

during winter months, manatees use the warm<br />

water springs in Crystal River to rest at night and<br />

often travel downriver to feed on Ruppia during<br />

the day (Reep & Bonde, 2006). Similar evidence<br />

of resting during certain times has been found for<br />

other manatee species. Observations of a solitary<br />

Amazonian manatee (T. inunguis) suggested an<br />

activity pattern with the individual sleeping for the<br />

first half of the night (Mukhametov et al., 1992). In<br />

areas with high instances of hunting, there is evidence<br />

that manatees (including the West African<br />

manatee, T. senegalensis) may have shifted to<br />

Resting Holes of the Antillean Manatee in Belize 63<br />

greater nocturnal activity and a more diurnal resting<br />

pattern (Rathbun et al., 1983; Reynolds et al.,<br />

1995; Powell & Kouadio, 2006).<br />

Manatees also appear to follow patterns in<br />

both resting times and resting sites. During low<br />

tides and the dry season in Venezuela and Costa<br />

Rica, they often take refuge in holes or channels<br />

in the rivers that range from 6 to 12 m in depth<br />

(O’Shea et al., 1988; Smethurst & Nietschmann,<br />

1999). Anecdotal observations suggest that manatees<br />

in Nicaragua rest in quiet, sheltered, deep<br />

water during most of the day, leaving to feed only<br />

during the night, early morning, and late afternoon<br />

(Jiménez, 2002). In Belize, resting behavior<br />

has been correlated with the presence of a “resting<br />

hole” (CSS & K. LaCommare, unpub. data),<br />

which has also been identified as a significant<br />

factor for predicting the presence of manatees at<br />

54 permanent sampling points in the Drowned<br />

Cayes during a daytime study (LaCommare et al.,<br />

2008). The objectives of this study were to determine<br />

what factors influence the use of previously<br />

identified resting holes by manatees and whether<br />

manatees use the resting holes more often during<br />

the day or at night.<br />

Materials and Methods<br />

Study Area<br />

The study was conducted over the summers of<br />

2005 and 2006 in the Drowned Cayes of Belize<br />

(located about 15 km east of Belize City at N <strong>17</strong>°<br />

28.0', W 88° 04.5'; Figure 1), which are a labyrinth<br />

of mangrove islands approximately 13 km<br />

long separated by lagoons, coves, and bogues<br />

(channels that separate the mangrove cayes both<br />

within and between ranges) (Ford, 1991). The<br />

ecosystem is marine with an average salinity of<br />

35 ppt, although a few low salinity sites have been<br />

identified (CSS, unpub. data). Extensive seagrass<br />

beds surround these islands, making them excellent<br />

foraging habitats for manatees.<br />

Characterization of Resting Holes<br />

For the purpose of this study, a resting hole is<br />

defined as an area where manatees have consistently<br />

been observed in a resting state of behavior.<br />

Twelve resting holes in the Drowned Cayes<br />

were chosen as a subset of 54 permanent scanpoints<br />

where manatees have been studied since<br />

2001 (LaCommare et al., 2008). The holes were<br />

characterized by depth using a HondexTM digital<br />

depth sounder (Forestry Suppliers), substrate<br />

(mud or sand), and vegetation (seagrass, algae,<br />

none). Water temperature (using a Taylor ® classic<br />

instant read pocket digital thermometer), salinity<br />

(using a WP-84 conductivity-TDS-temperature<br />

meter from TPS Pty), and water velocity (using


64 Bacchus et al.<br />

Figure 1. Map of Belize (inset) and the Drowned Cayes in relation to Belize City<br />

a General Oceanics mechanical flowmeter model<br />

2030R from Forestry Suppliers; speed range<br />

10 cm/s to 7.9 m/s) were measured at three points<br />

in the water column—surface, mid-water, and<br />

bottom—above each resting hole. Latitude and<br />

longitude were recorded above the deepest point<br />

of each resting hole with a GPS (Garmin ® Global<br />

Positioning System 12 Personal Navigator ® ). For<br />

comparison, these variables were measured at 20<br />

non-resting hole scan-points.


Diurnal and Nocturnal Scans<br />

Four of the 12 resting holes were selected for systematic<br />

diurnal and nocturnal scans during the summers<br />

of 2005 and 2006: B1Co, 30 CCo, 31 CLa,<br />

and 50 GiCr. Two 30-min scans focused around<br />

a fixed survey point were conducted at each site<br />

daily, once between 1000 and 1500 h, and again<br />

between 1900 and 0000 h (sunset occurred ~1900<br />

h). This survey method is referred to as a point<br />

scan (Self-Sullivan et al., 2003; LaCommare et al.,<br />

2008). The order in which sites were scanned was<br />

randomly chosen by the researcher each day. An<br />

8-m fiberglass boat with an outboard engine was<br />

used with observers standing in the boat during<br />

the scans. As the scan-point was approached, the<br />

boat engine was turned off, a long pole used to<br />

pull the boat to the scan-point, and the boat tied<br />

to the grounded pole. During each point scan, the<br />

research team continuously searched for manatees<br />

by performing 180° visual scans from different<br />

points on the boat and listening for breaths (for<br />

detailed point scan methods, see Bacchus, 2007;<br />

Self-Sullivan, 2008). Nocturnal scans were performed<br />

in the same manner as diurnal scans, with<br />

the addition of spotlights to help sight manatees.<br />

Data collected during each point scan included<br />

manatee sighting (a sighting was defined as detection<br />

of one or more manatees during the 30-min<br />

scan), sea and weather conditions, tidal state, sea<br />

surface temperature (SST), salinity, and air temperature.<br />

Data Analyses<br />

All statistics were run in SPSS, Version 13.0,<br />

with α = 0.05. Summary statistics were run for<br />

resting hole characterizations and data collected<br />

during the diurnal and nocturnal scans. The variable<br />

“depth” was adjusted to account for tides<br />

using predictions from a Belize City tide chart<br />

and visually verified via water level marks on the<br />

mangrove roots during each scan. The variable<br />

“water velocity” was calculated from 5-min rotation<br />

counts using the formula provided by General<br />

Oceanics:<br />

Resting Holes of the Antillean Manatee in Belize 65<br />

(1)<br />

where 26,873 is the rotor constant, and 999,999<br />

is the maximum number of rotations divided<br />

by 5 min and multiplied by 60 s to obtain cm/s.<br />

T-tests were used to test for differences in depth<br />

and water velocity between areas with and areas<br />

without resting holes. Water velocity was ranked to<br />

adjust for unequal variances to meet assumptions<br />

of normality. T-tests were also run to analyze SST<br />

and salinity between the two years, and between<br />

day and night. Log likelihood ratio tests (G 2 ) were<br />

run to test for variance in the number of manatee<br />

sightings between the two years and between day<br />

and night. A step-wise, logistic regression was<br />

used to detect any associations between specific<br />

habitat factors and manatee sightings. Factors<br />

included location, tide, salinity, SST, day/night,<br />

and year.<br />

Results<br />

Characterization of Resting Holes<br />

There was a significant difference in water depth<br />

between resting holes and scan-points without<br />

resting holes (t(15) = 4.541, p < 0.001) with a mean<br />

difference of 1.5 ± 0.32 m. Mean water depth for<br />

resting holes was 3.5 ± 0.30 m (n = 12; range: 2.0<br />

to 5.2 m) compared with non-resting hole areas<br />

with a mean depth of 2.0 ± 0.12 m (n = 20; range:<br />

1.4 to 3.3 m) (Table 1).<br />

Mean surface water velocity for resting holes<br />

was 0.89 ± 0.51 cm/s (n = 10; range: 0 to 5.20<br />

cm/s). Non-resting hole areas had a mean surface<br />

water velocity of 4.26 ± 1.14 cm/s (n = 20; range:<br />

0.01 to <strong>17</strong>.12 cm/s) (Table 1). These represent a<br />

significant difference in surface water velocity<br />

between resting holes and areas without resting<br />

holes (t(28) = -2.880, p = 0.008). All of the holes<br />

characterized had little or no bottom vegetation.<br />

The substrate in and around the resting holes<br />

tended to be a silty mixture of mud and sand,<br />

which was easily compressed and resuspended.<br />

Sparse to abundant vegetation was found proximal<br />

to resting holes, usually consisting of shoal<br />

grass (Halodule sp.), macro algae (Halimeda and<br />

Penicillus sp.), and/or turtle grass. One of the resting<br />

holes also had a visible “manatee highway” to<br />

and from the hole where manatees swim or paddle<br />

along the bottom with their forelimbs and/or torso<br />

in contact with the seafloor. This deeper, narrow<br />

channel started at the entrance of the small cove<br />

containing the resting hole near the mangroves<br />

and followed a straight path to the deepest part of<br />

the resting hole.<br />

Diurnal and Nocturnal Scans<br />

A total of 168 scans were conducted at four resting<br />

hole sites during the summer of 2005 (n = 81)<br />

and the summer of 2006 (n = 87). Manatees were<br />

sighted during 39 scans—26 times in 2005 and 13<br />

times in 2006 (Figure 2)—representing a significant<br />

difference in the number of sightings at these<br />

four sites between years (G 2 (1) = 7.01, p < 0.01).<br />

Table 2 shows that SST and salinity were significantly<br />

different for these four points between<br />

years (SST: t(158) = 3.292, p = 0.001; salinity: t(166) =<br />

20.629, p < 0.001). There was no significant difference<br />

in air temperature between years.


66 Bacchus et al.<br />

Table 1. Descriptive statistics of environmental characteristics of areas without and with resting holes; means are reported<br />

with ± 1 SE.<br />

There were 107 day scans and 61 night scans<br />

over both summers, with 34 day sightings<br />

and 5 night sightings (Figure 3), resulting in a<br />

significant difference in the number of manatee<br />

sightings between day and night (G 2 (1) = 13.68, p <<br />

0.001). Table 3 shows that both sea surface and air<br />

temperatures were significantly different between<br />

day and night, although surface salinity was not<br />

significantly different between day and night.<br />

A forward, step-wise logistic regression was conducted<br />

to determine which independent variables<br />

Non-resting hole sites Resting hole sites<br />

N Min Max Mean ± SE N Min Max Mean ± SE Significance<br />

Depth (m) 20 1.4 3.3 2.0 ± 0.12 12 2.0 5.2 3.5 ± 0.30


Figure 3. The number of scans (white bars) and sightings<br />

(black bars) during diurnal (left) and nocturnal (right) scans<br />

for each location<br />

Characterization of Resting Holes<br />

Sites with resting holes were located in dead end<br />

bogues, lagoons, and coves close to the mangroves<br />

with little current and calm sea state. Water<br />

depth was greater and surface velocity was lower<br />

at resting holes than at scan-points without resting<br />

holes. However, comparison of water velocity<br />

between resting holes could not be made because<br />

measurements within resting holes were all less<br />

than 10 cm/s, which was the lower threshold for<br />

the flowmeter. Slow current and calm waters may<br />

enable manatees to rest without exerting energy<br />

to hold their position. These results support previous<br />

studies that reported manatees are found in<br />

areas sheltered from high currents, surf, and wind<br />

(Hartman, 1979; Powell et al., 1981; Lefebvre<br />

et al., 2001; Jiménez, 2005; Olivera-Gomez &<br />

Mellink, 2005).<br />

Although there was no seagrass growing in<br />

the resting holes themselves, many seagrass beds<br />

were nearby. This would facilitate a short distance<br />

of travel for manatees between resting sites and<br />

feeding sites in calm waters. Given their large<br />

body mass and dependence on a low-energy, lowprotein<br />

diet, manatees must spend a large proportion<br />

of their time feeding to meet metabolic<br />

requirements. Manatees tend to feed from 5 to<br />

8 h/d, consuming about 7% of their body weight in<br />

wet vegetation (Hartman, 1979; Bengtson, 1983;<br />

Resting Holes of the Antillean Manatee in Belize 67<br />

Etheridge et al., 1985). Manatees may choose foraging<br />

strategies that allow them to maximize food<br />

intake while minimizing energy output. Similar<br />

studies in the Drowned Cayes indicated that sightings<br />

of manatees were most probable near resting<br />

holes and seagrass beds (LaCommare et al., 2008).<br />

The more familiar an animal is with locations of<br />

essential resources, the more efficient it will be in<br />

energy acquisition since this reduces searching<br />

time (Deutsch et al., 2003).<br />

Manatees were frequently encountered at three<br />

of the four previously identified resting holes<br />

within the Drowned Cayes. All four holes were<br />

located in protected coves or lagoons with minimal<br />

current, making it likely that manatees choose<br />

these areas because of the quiet conditions and<br />

create, or at least maintain, resting holes at these<br />

sites. Although previously categorized as a resting<br />

hole (LaCommare et al., 2008), the scan-point<br />

B1Co was likely not being used as a resting site in<br />

the summers of 2005 and 2006; it had a uniform<br />

1.5 m depth, with patches of shoal grass and no<br />

depression. However, there were many feeding<br />

scars and forelimb marks where manatees had dug<br />

into the mud to eat seagrass roots and rhizomes.<br />

Only one manatee was observed resting at this<br />

site in the summer of 2005, and no manatees were<br />

observed resting in the cove during the summer<br />

of 2006. This may indicate that manatee use of<br />

particular resting holes changes over time.<br />

Although manatees frequently used known<br />

resting holes, there was no evidence that manatees<br />

intentionally excavated or maintained the holes.<br />

Our current hypothesis is that they make use of<br />

natural depressions at these sites. The substrate<br />

in the holes is primarily uncompacted silt with<br />

high interstitial space, which is easily stirred-up<br />

as manatees rise off the bottom to surface for<br />

breaths. By resting regularly in the same place,<br />

the presence and vertical movement of a manatee<br />

during breathing cycles could easily deepen and<br />

maintain the depression.<br />

However, we do not yet know if individual<br />

manatees return to preferred holes exclusively<br />

or if resting hole use is indiscriminate. Also, it is<br />

unknown how manatees learn about the presence<br />

Table 3. Descriptive statistics for sea surface temperature, surface salinity, and air temperature during day and night scans<br />

of four locations<br />

Day Night<br />

N Min Max Mean ± SE N Min Max Mean ± SE Significance<br />

Sea surface temperature (°C) 106 27.2 34.3 30.9 ± 0.12 61 28.6 31.8 30.4 ± 0.10 0.001<br />

Surface salinity (ppt) 107 13.0 39.0 34.0 ± 0.30 61 28.0 37.0 34.0 ± 0.30 NS<br />

Air temperature (°C) 107 26.6 33.4 31.5 ± 0.12 61 27.4 32.0 29.4 ± 0.08


68 Bacchus et al.<br />

of these specific resting holes. In Florida, individual<br />

manatees show strong site fidelity to warmseason<br />

and winter ranges where manatee mothercalf<br />

pairs have been tracked remotely over time<br />

(Deutsch et al., 2003). Use of specific sites, initially<br />

with their mothers and then as independent<br />

subadults, has demonstrated natal philopatry to<br />

specific warm-season and winter ranges as well<br />

as to migratory patterns and travel routes among<br />

sites. Similarly, Antillean manatees may learn of<br />

specific resting holes as dependent calves and<br />

return to these same areas as adults.<br />

Diurnal and Nocturnal Use of Resting Holes<br />

There was a significant difference in the number<br />

of sightings of manatees between day and night.<br />

Although spotlights were used to increase the<br />

probability of detection during night scans, we<br />

acknowledge the increased potential for observer<br />

error during nocturnal scans. However, this potential<br />

was further reduced by increased vigilance<br />

during night scans, both visually and aurally. To<br />

increase successful detection and maintain consistency<br />

in effort, the same researchers and assistants<br />

were used for all night scans during this study.<br />

Despite concerns over detectability, these data<br />

suggest that manatees used resting holes more<br />

frequently during the day than at night in the<br />

Drowned Cayes. These results are consistent with<br />

observations in Florida where manatees leave<br />

the warm water refuge of Blue Spring in the late<br />

winter afternoons to feed all night, returning to the<br />

refuge in the morning (Bengtson, 1981; Rathbun<br />

et al., 1990). However, the animals abandoned<br />

the diurnal resting behavior in Blue Spring when<br />

water temperatures warmed in the spring. If use<br />

of resting sites is driven more by resources than<br />

a daily activity pattern, perhaps manatees use the<br />

resting holes in the current study area during the<br />

day when human activity on the water is high,<br />

then leave at night to feed in nearby seagrass beds<br />

or travel to freshwater sites. Interviews with local<br />

people indicated a similar pattern in Nicaragua<br />

where manatees rest in deep, quiet waters during<br />

the day and move to shallow waters to feed during<br />

late afternoon, night, and early morning, presumably<br />

to avoid hunting predation (Jiménez, 2002).<br />

Another possible reason for different resting<br />

hole uses during the day and night is the need for<br />

fresh water. Behavioral, ecological, and physiological<br />

evidence suggest that manatees require<br />

fresh or low salinity water for osmoregulation<br />

when living in a marine environment and feeding<br />

exclusively on seagrasses (Hartman, 1979; Powell<br />

& Rathbun, 1984; Ortiz et al., 1998, 1999). A lack<br />

of fresh water for extended periods leads to dehydration,<br />

although manatees may be able to osmoregulate<br />

via oxidation of fat for short periods of<br />

time (Ortiz et al., 1999). In Puerto Rico, where<br />

manatees are primarily observed in marine habitats,<br />

Powell et al. (1981) noted that 85.8% of<br />

manatee sightings were within 5 km of freshwater<br />

sources. Those authors also reported anecdotes<br />

by local fisherman indicating that manatees often<br />

visited nearby rivers to drink water. In aerial surveys<br />

in Belize, a higher number of manatees have<br />

been observed in near-shore habitat (e.g., estuary,<br />

lagoon, and river) than in cay habitat, with one<br />

of the highest numbers recorded from the Belize<br />

River (Auil, 2004). Also, more manatees were<br />

sighted in near-shore habitat during the dry season<br />

(November to May), and more manatees were<br />

sighted in the cay habitat during the wet season<br />

(May/June to November) (Auil, 2004).<br />

Similar to the situation in Ten Thousand Islands,<br />

Florida (Butler et al., 2003; J. P. Reid, pers.<br />

comm.), manatees in the Drowned Cayes travel to<br />

the Belize River, the closest source of substantial<br />

fresh water (about 12 to 15 km away) during the dry<br />

season (Auil, 2004). Manatees may also be traveling<br />

to the Belize River on a shorter time frame.<br />

Since boat traffic, both on the river and between<br />

the river and the cayes, is much higher during the<br />

day, manatees that travel from the Drowned Cayes<br />

to the river at night may reduce the probability<br />

of being struck by boats. This hypothesis is supported<br />

by studies in Florida and Costa Rica, where<br />

manatees have been known to avoid areas of high<br />

boat traffic, changing their behavioral state and<br />

moving out of a geographical area in response to<br />

approaching vessels (Buckingham et al., 1999;<br />

Smethurst & Nietschmann, 1999; Miksis-Olds,<br />

2006).<br />

There was a significant difference in sightings<br />

between the two summers with twice as<br />

many manatees seen in the summer of 2005 than<br />

in the summer of 2006. The reasons for this are<br />

unclear, but longer-term data for all 54 scanpoints<br />

detected no significant difference in the<br />

probability of encountering manatees over the<br />

period of 2001 to 2004, even as a result of season<br />

(Self-Sullivan, 2008). There was a significant<br />

difference in water temperature and salinity for<br />

the four scan locations between these two summers,<br />

with 2005 being warmer and more saline.<br />

However, the difference in temperature (0.5° C),<br />

although significantly different between years, is<br />

relatively small and probably not significant to<br />

manatees given what is known about their range.<br />

A similar situation occurred with day and night<br />

SST with only a 0.5° C difference. Although manatees<br />

can sense water temperature and salinity differences<br />

(Deutsch et al., 2003), the differences in<br />

the temperature readings in this study seem small<br />

enough to insignificantly affect manatee habitat<br />

choices. Salinity was 36 ppt in 2005 and 32


in 2006. Although this difference is fairly small,<br />

this may indicate a slightly more rainy summer<br />

season, which might have some impact on the<br />

probability of manatee presence at the four resting<br />

holes studied. Even though manatees were<br />

not seen as often in the summer of 2006, this does<br />

not necessarily mean that they used the Drowned<br />

Cayes habitat less frequently but rather that they<br />

were not detected as often at the particular resting<br />

holes being studied.<br />

West Indian manatee populations must deal<br />

with environmental changes on a daily (e.g.,<br />

tides, currents, weather, human activity levels),<br />

seasonal (e.g., rainfall, temperature), and generational<br />

basis as human development continues to<br />

dramatically alter their habitats throughout the<br />

Caribbean (Deutsch et al., 2007). Although little<br />

is known about habitat use in the population over<br />

long periods of time, changes in habitat use by<br />

manatees have been observed at Swallow Caye<br />

in the Drowned Cayes of Belize, presumably as<br />

a result of increased human disturbance. When<br />

tour operators began and then ended a "swimwith"<br />

program at Swallow Caye, manatee use of<br />

the resting hole decreased during the program but<br />

increased again when the program was discontinued<br />

(Self-Sullivan, 2008). For generations of<br />

manatees, the undeveloped Drowned Cayes have<br />

provided a safe haven within a short distance of<br />

essential fresh water from the Belize River.<br />

Between 2001 and 2004, there was an exponential<br />

increase in cruise ship tourism in Belize.<br />

In 2001, there were 48 cruise ships with 48,116<br />

passengers; in 2002, there were 200 ships with<br />

319,690 passengers, representing a 584% growth;<br />

in 2003, there were 315 ships with 575,196 passengers,<br />

an 80% growth from 2002; and in 2004,<br />

there were 406 ships with 851,436 passengers,<br />

which is a 49% growth from 2003 (Belize Tourism<br />

Board, 2006). Although cruise ship numbers have<br />

leveled-off since 2005, the impact of cruise ship<br />

tourism remains above the 2003 level with a continued<br />

increase in development and human disturbance<br />

of manatee habitat in the Drowned Cayes<br />

area. Manatee mortality rate near Belize City is<br />

relatively high (Auil, 2004). This might be due to<br />

the increase in boat traffic as passengers are ferried<br />

from the ships by tender boats and are taken<br />

on tours to the reef, around the Drowned Cayes,<br />

and to small sandy cayes in the area (Self-Sullivan,<br />

2008). Also, there has been an increase in development<br />

of some of the other cayes in the Drowned<br />

Cayes, causing increases in boat traffic in these<br />

areas with the potential to cause siltation that may<br />

destroy the seagrass beds that the manatees feed<br />

on.<br />

The implications of our findings should be taken<br />

into consideration by conservation managers who<br />

Resting Holes of the Antillean Manatee in Belize 69<br />

are concerned with development in the Drowned<br />

Cayes. Conservation strategies designed to protect<br />

manatees should include preservation of<br />

low-energy, secluded areas where manatees can<br />

continue to maintain resting holes near seagrass<br />

beds. Results of the current study not only increase<br />

understanding of manatee resting habitats but provide<br />

necessary data for decisionmakers and wildlife<br />

managers to establish limits of acceptable change<br />

within the Drowned Cayes area and other strategies<br />

for the conservation of manatee habitat in Belize.<br />

Acknowledgments<br />

This work was supported by a Marine Research<br />

Group (LLU) grant to M-LB, funding from Dr.<br />

Robert Ford in the Department of Social Work and<br />

Social Ecology (LLU), the Earthwatch Institute,<br />

and the Hugh Parkey Foundation for Marine<br />

Awareness and Education. We thank our Belizean<br />

field assistants and boat drivers, Gilroy Robinson<br />

and Dorian Alvarez, as well as field assistants<br />

Leigh Bird, Shauna King, Haydée Domínguez,<br />

Karen Petkau, Ryan Roland, Austin Bacchus,<br />

Coralie Lallemand, and the volunteers from the<br />

Earthwatch Teams 2, 3, and 4 in 2005 and 2006.<br />

We are also grateful to multiple colleagues at<br />

Loma Linda University; Dr. William Hayes and<br />

Dr. Zia Nisani who helped with statistics; and<br />

Daniel Gonzalez who helped in the field. Thanks<br />

to Rick Ware for providing the Belize City tide<br />

tables. Our research was carried out in compliance<br />

with the laws of the United States and Belize under<br />

MNREI Forest Department Permits, Ref. Nos.<br />

CD/60/3/05(29) and CD/60/3/06(32), and LLU<br />

IACUC Protocol #86006. This is contribution<br />

Number 8 of the Marine Research Group (LLU).<br />

Literature Cited<br />

Auil, N. E. (2004). Abundance and distribution trends of<br />

the West Indian manatee in the coastal zone of Belize:<br />

Implications for conservation. Master’s thesis, Texas<br />

A&M University, College Station. 83 pp.<br />

Axis-Arroyo, J., Morales-Vela, B., Torruco-Gomez, D., &<br />

Vega-Cendejas, M. E. (1998). Variables asociadas con el<br />

uso de habitat del manatí del Caribe (Trichechus manatus),<br />

en Quintana Roo, México (Mammalia). Revista de<br />

Biologia Tropical, 46, 791-803.<br />

Bacchus, M-L. C. (2007). Characterization of resting holes<br />

and use by the Antillean manatee (Trichechus manatus<br />

manatus). Master’s thesis, Loma Linda University, CA.<br />

84 pp.<br />

Belize Tourism Board. (2006). Cruise arrivals by month.<br />

Retrieved December 2006 from www.belizetourism.org.<br />

Bengtson, J. L. (1981). Ecology of manatees (Trichechus<br />

manatus) in the St. Johns River, Florida. Ph.D. dissertation,<br />

University of Minnesota, Minneapolis. 126 pp.


70 Bacchus et al.<br />

Bengtson, J. L. (1983). Estimating food consumption of<br />

free-ranging manatees in Florida. Journal of Wildlife<br />

Management, 47, 1186-1192.<br />

Buckingham, C. A., Lefebvre, L. W., Schaefer, J. M., &<br />

Kochman, H. I. (1999). Manatee response to boating<br />

activity in a thermal refuge. Wildlife Society Bulletin,<br />

27, 514-522.<br />

Butler, S. M., Reid, J. P., & Stith, B. M. (2003). Detailed<br />

movements and habitat use patterns of radio tagged<br />

manatees in the western Everglades. Poster presentation<br />

at the 15th Biennial Conference on the Biology of<br />

Marine Mammals. Greensboro, NC.<br />

Deutsch, C. J., Self-Sullivan, C., & Mignucci-Giannoni,<br />

A. A. (2007). Trichechus manatus. In 2007 IUCN<br />

red list of threatened species. Gland, Switzerland:<br />

<strong>International</strong> Union for Conservation of Nature and<br />

Natural Resources.<br />

Deutsch, C. J., Reid, J. P., Bonde, R. K., Easton, D. E.,<br />

Kochman, H. I., & O’Shea, T. J. (2003). Seasonal movements,<br />

migratory behavior, and site fidelity of West<br />

Indian manatees along the Atlantic coast of the United<br />

States. Wildlife Monographs, 151, 1-77.<br />

Domning, D. P., & Hayek, L. C. (1986). Interspecific<br />

and intraspecific morphological variation in manatees<br />

(Sirenia: Trichechus). Marine Mammal Science, 2,<br />

87-144.<br />

Etheridge, K., Rathbun, G. B., Powell, J. A., & Kochman,<br />

H. I. (1985). Consumption of aquatic plants by the West<br />

Indian manatee. Journal of Aquatic Plant Management,<br />

23, 21-25.<br />

Ford, R. E. (1991). Toponymic generics, environment and<br />

culture history in pre-independence Belize. Names, 39,<br />

1-25.<br />

Hartman, D. S. (1971). Behavior and ecology of the Florida<br />

manatee, Trichechus manatus latirostris (Harlan), at<br />

Crystal River, Citrus County. Ph.D. dissertation, Cornell<br />

University, Ithaca, NY. 285 pp.<br />

Hartman, D. S. (1979). Ecology and behavior of the manatee<br />

(Trichechus manatus) in Florida. American Society<br />

of Mammalogists, Special Publication 5.<br />

Jiménez, I. (2002). Heavy poaching in prime habitat:<br />

The conservation status of the West Indian manatee in<br />

Nicaragua. Oryx, 36, 272-278.<br />

Jiménez, I. (2005). Development of predictive models to<br />

explain the distribution of the West Indian manatee<br />

Trichechus manatus in tropical watercourses. Biological<br />

Conservation, 125, 491-503.<br />

Kinnaird, M. F. (1985). Aerial census of manatees in northeastern<br />

Florida. Biological Conservation, 32, 59-79.<br />

LaCommare, K. S., Self-Sullivan, C., & Brault, S. (2008).<br />

Distribution and habitat use of Antillean manatees<br />

(Trichechus manatus manatus) in the Drowned Cayes<br />

area of Belize, Central America. Aquatic Mammals,<br />

34(1), 35-43.<br />

Lefebvre, L. W., Marmontel, M., Reid, J. P., Rathbun, G. B.,<br />

& Domning, D. P. (2001). Status and biogeography of<br />

the West Indian manatee. In C. A. Wood & F. E. Sergile<br />

(Eds.), Biogeography of the West Indies: Patterns and<br />

perspectives (2nd ed.) (pp. 425-474). Boca Raton, FL:<br />

CRC Press.<br />

Miksis-Olds, J. L. (2006). Manatee response to environmental<br />

noise. Ph.D. dissertation, University of Rhode<br />

Island, Narragansett. 228 pp.<br />

Mukhametov, L. M., Lyamin, O. I., Chetyrbok, I. S.,<br />

Vassilyev, A. A., & Diaz, R. P. (1992). Sleep in an<br />

Amazonian manatee, Trichechus inunguis. Experientia,<br />

48, 4<strong>17</strong>-419.<br />

Olivera-Gomez, L. D., & Mellink, E. (2005). Distribution<br />

of the Antillean manatee (Trichechus manatus manatus)<br />

as a function of habitat characteristics, in Bahia de<br />

Chetumal, Mexico. Biological Conservation, 121, 127-<br />

133.<br />

Ortiz, R. M., Worthy, G. A. J., & Byers, F. M. (1999).<br />

Estimation of water turnover rates of captive West Indian<br />

manatees (Trichechus manatus) held in fresh and salt<br />

water. Journal of Experimental Biology, 202, 33-38.<br />

Ortiz, R. M., Worthy, G. A. J., & MacKenzie, D. S. (1998).<br />

Osmoregulation in wild and captive West Indian manatees<br />

(Trichechus manatus). Physiological Zoology, 71,<br />

449-457.<br />

O’Shea, T. J., Correa-Viana, M., Ludlow, M. E., & Robinson,<br />

J. G. (1988). Distribution, status, and traditional significance<br />

of the West Indian manatee Trichechus manatus in<br />

Venezuela. Biological Conservation, 46, 281-301.<br />

Packard, J. M., & Wetterqvist, O. F. (1986). Evaluation of<br />

manatee habitat systems on the northwestern Florida<br />

coast. Coastal Zone Management Journal, 14, 279-310.<br />

Powell, J. A., & Kouadio, A. (2006). Trichechus senegalensis.<br />

In 2007 IUCN red list of threatened species. Gland,<br />

Switzerland: <strong>International</strong> Union for Conservation of<br />

Nature and Natural Resources.<br />

Powell, J. A., & Rathbun, G. B. (1984). Distribution and<br />

abundance of manatees along the northern coast of the<br />

gulf of Mexico. Northeast Gulf Science, 7, 1-28.<br />

Powell, J. A., Belitsky, D. W., & Rathbun, G. B. (1981).<br />

Status of the West Indian manatee (Trichechus manatus)<br />

in Puerto Rico. Journal of Mammalogy, 62, 642-646.<br />

Provancha, J. A., & Hall, C. R. (1991). Observations of<br />

associations between seagrass beds and manatees in east<br />

central Florida. Florida Scientist, 54, 87-98.<br />

Ralph, C. L., Young, S., Gettinger, R., & O’Shea, T. J.<br />

(1985). Does the manatee have a pineal body? Acta<br />

Zoologica, 66, 55-60.<br />

Rathbun, G. B., Powell, J. A., & Cruz, G. (1983). Status<br />

of the West Indian manatee in Honduras. Biological<br />

Conservation, 26, 301-308.<br />

Rathbun, G. B., Reid, J. P., & Carowan, G. (1990).<br />

Distribution and movement patterns of manatees<br />

(Trichechus manatus) in northwestern peninsular<br />

Florida. Florida Marine Research Publications, 48,<br />

1-33.<br />

Reep, R. L., & Bonde, R. K. (2006). The Florida manatee:<br />

Biology and conservation. Gainesville: University Press<br />

of Florida. 189 pp.<br />

Reynolds, J. E., Szelistowski, W. A., & Leon, M. A. (1995).<br />

Status and conservation of manatees Trichechus manatus


manatus in Costa Rica. Biological Conservation, 71,<br />

193-196.<br />

Self-Sullivan, C. (2008). Conservation of Antillean manatees<br />

in the Drowned Cayes area of Belize. Ph.D. dissertation,<br />

Texas A&M University, College Station. 129 pp.<br />

Self-Sullivan, C., Smith, G. W., Packard, J. M., &<br />

LaCommare, K. S. (2003). Seasonal occurrence of male<br />

Antillean manatees (Trichechus manatus manatus) on<br />

the Belize Barrier Reef. Aquatic Mammals, 29(3), 342-<br />

354.<br />

Smethurst, D., & Nietschmann, B. (1999). The distribution<br />

of manatees (Trichechus manatus) in the coastal<br />

waterways of Tortuguero, Costa Rica. Biological<br />

Conservation, 89, 267-274.<br />

Resting Holes of the Antillean Manatee in Belize 71


J Coast Conserv (2011) 15:573<strong>–</strong>583<br />

DOI 10.1007/s11852-011-0146-3<br />

Aerial surveys of manatees (Trichechus manatus) in Lee<br />

County, Florida, provide insights regarding manatee<br />

abundance and real time information for managers<br />

and enforcement officers<br />

Deirdre J. Semeyn & Carolyn C. Cush &<br />

Kerri M. Scolardi & Jennifer Hebert &<br />

Justin D. McBride & Denis Grealish & John E. Reynolds<br />

Received: 8 October 2010 /Accepted: 7 January 2011 /Published online: 28 January 2011<br />

# Springer Science+Business Media B.V. 2011<br />

Abstract Conservation and management of the endangered<br />

Florida manatee is often centered on reducing mortality<br />

caused by watercraft collisions. Lee County, Florida, has<br />

led the state in watercraft-related mortality for eight of the<br />

last 10 years. This county is of particular concern as it<br />

contains important habitat for manatees, including extensive<br />

feeding grounds and an artificial warm-water refuge<br />

where more than 900 manatees have been recorded on a<br />

single day. Distributional aerial surveys were conducted<br />

from April 2007 through April 2009 over Lee County<br />

waters. Surveys yielded higher numbers of manatees than<br />

D. J. Semeyn<br />

5507 S. Bernie St.,<br />

Tampa, FL 33611, USA<br />

C. C. Cush : K. M. Scolardi : J. E. Reynolds (*)<br />

Mote Marine Laboratory,<br />

1600 Ken Thompson Parkway,<br />

Sarasota, FL 34236, USA<br />

e-mail: reynolds@mote.org<br />

J. Hebert<br />

2740 Chariton Street,<br />

Oakton, VA 22124, USA<br />

J. D. McBride<br />

Lee County Division of Natural Resources Marine Program,<br />

1500 Monroe Street,<br />

Fort Myers, FL 33902, USA<br />

D. Grealish<br />

Florida Fish and Wildlife Conservation Commission,<br />

Division of Law Enforcement,<br />

Edwards Drive,<br />

Fort Myers, FL 33901, USA<br />

previously observed in this area. Using GIS methodology,<br />

kernel density analysis illustrated seasonal changes in<br />

distribution patterns and highlighted areas where manatees<br />

were most densely clustered. For example, during summer<br />

months, manatees were widely distributed throughout the<br />

survey area, with high-density areas associated with<br />

seagrass beds. During winter months, manatees were<br />

densely clustered at warm-water sites and over feeding<br />

grounds within close distance of these sites. These seasonal<br />

distribution patterns coincide well with speed zone designations.<br />

Counts and distributions of manatees were made<br />

available, almost immediately if necessary, to local marine<br />

law enforcement in an attempt to focus resources toward<br />

reducing manatee-watercraft collisions. Future studies<br />

should implement similar communication strategies to<br />

improve conservation efforts.<br />

Keywords Florida manatee . Aerial survey . Kernel density<br />

analysis . Lee County . Conservation . Management<br />

Introduction<br />

Over the past 40 years, aerial surveys have been used to<br />

monitor and assess number and distribution of Florida<br />

manatees, Trichechus manatus latirostris (e.g., Craig and<br />

Reynolds 2004; Wright et al. 2002). Aerial survey data are<br />

one of the primary sources of information used to create<br />

state and federal manatee sanctuaries and boat speed<br />

regulatory zones, as well as individual county manatee<br />

management plans (Laist and Shaw 2006).<br />

Listed as endangered under the Endangered Species Act of<br />

1973, the Florida manatee continues to face anthropogenic


574 D. J. Semeyn et al.<br />

and natural hazards that threaten the long-term viability of the<br />

species. Human related threats contributing to manatee<br />

mortality include watercraft collisions and coastal development<br />

leading to habitat loss and degradation (e.g. effects on<br />

warm-water availability). Naturally occurring events such as<br />

harmful algal blooms, intense coastal storms, and prolonged<br />

exposure to cold temperatures are also major causes of<br />

mortality in manatees (Ackerman et al. 1995; Langtimmand<br />

Beck 2003). The subpopulation of manatees in southwestern<br />

Florida is affected by all of these anthropogenic and natural<br />

factors, leading to the highest annual mortality rate (Florida<br />

Fish and Wildlife Conservation Commission 2010) and<br />

lowest estimated adult survival rates for manatees in the<br />

state (Langtimm et al. 2004).<br />

Lee County is often the focus of scrutiny with regard to<br />

management of manatees in southwestern Florida. This<br />

county contains critical habitat for manatees with over 590<br />

miles of natural shoreline (USA Cities 2009), extensive<br />

seagrass beds, and three important warm-water refuges:<br />

Florida Power & Light Company power plant (FPL) on the<br />

Orange River, the dredged canals of Matlacha Isles, and 10<br />

Mile Canal off of Estero Bay. It is also home to nearly<br />

600,000 people, and 50,464 total registered vessels, 48,853<br />

of which are recreational (United States Census Bureau<br />

2009; Florida Fish and Wildlife Conservation Commission<br />

2008). Consequently, manatee and human use of waterways<br />

overlap, creating increased potential for watercraft related<br />

collisions and mortality.<br />

The Florida Fish and Wildlife Conservation Commission<br />

(FWC) and U.S. Fish and Wildlife Service (USFWS) have<br />

developed and implemented boat speed and boat access<br />

regulations throughout the county in an attempt to reduce<br />

watercraft mortality. Seven types of regulatory zones exist<br />

throughout Lee County, although it is mainly comprised of three:<br />

1) Slow speed all year;<br />

2) Slow speed April 1-November 15; 25 mph rest of year;<br />

and<br />

3) 25 mph within marked channel.<br />

Even with these regulatory zones, Lee County surpasses<br />

almost all other counties in terms of annual human-related<br />

mortality of manatees, with an average of 15 per year over the<br />

last 10 years (Florida Fish and Wildlife Conservation<br />

Commission 2010). The Potential Biological Removal level<br />

(PBR) determined for the statewide population of manatees<br />

is 12 individuals (U.S. Fish and Wildlife Service 2009). 1<br />

Thus, human-related mortality in this one county exceeds the<br />

current (i.e. 2010) maximum number of manatees that may<br />

1 NOTE that this figure was calculated prior to the extremely high<br />

aerial survey counts of January 2010. As a consequence of there being<br />

more manatees than was previously believed, the PBR value of 12<br />

manatees is likely to be adjusted upward.<br />

be removed from the entire statewide population by such<br />

causes each year (U.S. Fish and Wildlife Service 2009).<br />

The original objective of this study was to conduct aerial<br />

surveys in order to provide current distributional data to Lee<br />

County managers for use in reviews of permits for new or<br />

modified marine facilities or other developments. This county<br />

is one of the 13 “key counties” in Florida, which are mandated<br />

by the State to develop and implement a Manatee Protection<br />

Plan (MPP). A major component of each MMP is the Marine<br />

Facility Siting Element (MFSE). The MFSE component of the<br />

Lee County MPP creates a slip-to-shoreline ratio for existing<br />

and proposed projects, based upon each project’s projected<br />

risk to manatees. The MFSE utilizes a series of input<br />

components including manatee mortality and distribution<br />

data. Whereas mortality data are updated each calendar year,<br />

until the completion of this project, Lee County resource<br />

managers had used distribution data collected in 1997<strong>–</strong>1998<br />

to evaluate marine facilities proposals. As the study progressed,<br />

we realized that managers could utilize survey data<br />

immediately to enhance, and later to evaluate, two other<br />

features of the MPP: law enforcement coordination and sitespecific<br />

boat speed zones (mentioned above).<br />

This study was not designed to provide a statistically robust<br />

estimate of manatee abundance in Lee County. However, it<br />

does provide an approach for monitoring spatial and temporal<br />

changes in distribution and relative abundance of manatees in<br />

order to formulate effective and proactive management and<br />

enforcement for this highly protected endangered species.<br />

Methods<br />

Aerial survey<br />

Aerial surveys were selected as an efficient and cost effective<br />

method to collect information regarding number and distribution<br />

of manatees (Reynolds et al. 2010). Lee County<br />

encompasses 803.6 square miles, and is criss-crossed with<br />

rivers, creeks, islands and canals (United States Census<br />

Bureau 2009). To prevent observer fatigue and ensure the<br />

entire county was surveyed in optimal light and weather<br />

conditions, the survey area was divided into northern and<br />

southern sections. From April 2007 through April 2009, two<br />

simultaneous surveys were flown twice each month; one<br />

covered waters of the northern county, and the other surveyed<br />

waters of the southern portion. Each route followed a<br />

racetrack pattern along shorelines and over open water. With<br />

the exception of the Gulf of Mexico, the entirety of Lee<br />

County’s waterways was surveyed as well as five miles<br />

beyond the county boundary where a water body continued<br />

into Charlotte, Collier, or Hendry Counties. Two high-winged<br />

aircraft (Cessna <strong>17</strong>2 or 182) were used. Each plane flew at an<br />

altitude of 250 m and a speed of 80<strong>–</strong>90 knots.


Aerial surveys of manatees (Trichechus manatus) in Lee County, Florida 575<br />

A Garmin GPSmap 76 recorded the survey path in each<br />

aircraft. Observers recorded the number of manatees (adults<br />

and calves), location, habitat and behavior for each sighting.<br />

Manatees within close proximity of each other and displaying<br />

similar behavior were grouped together as one sighting. Wind<br />

speed and direction were recorded based on local airport<br />

conditions. Other environmental conditions such as water<br />

clarity, percent cloud cover, and water surface conditions were<br />

estimated and recorded by the surveyor. To ensure that water<br />

clarity and visibility were acceptable for efficient spotting of<br />

manatees, surveys were terminated when sustained winds<br />

exceeded 15 knots or surface conditions scored four or higher<br />

on the Beaufort Wind Scale.<br />

Spatial analysis<br />

Kernel density estimation calculates home range and animal<br />

density by estimating the probability of locating an<br />

individual at a specific place and time (Horne and Garton<br />

2006; Worton 1995). The method creates a “kernel” or<br />

probability density over each point in a dataset (Seaman<br />

and Powell 1996). The density estimate is calculated by<br />

averaging the densities of all kernels that overlap each point.<br />

Kernel density acts as a decay-function with high estimates in<br />

locations with numerous points (sightings), and low densities<br />

in locations with few points (Seaman and Powell 1996).<br />

Using ArcGIS 9.3 (ESRI), survey data were exported as<br />

point shapefiles, and the Spatial Analyst Extension was used to<br />

calculate kernel densities for the total number of manatees per<br />

time period in the survey area. To reflect seasonal shifts,<br />

consideration of manatee distribution and habitat use is often<br />

divided into two periods: winter (November 15<strong>–</strong>March 15) and<br />

non-winter (March 16<strong>–</strong>November 14). These two seasonal<br />

designations were used to create two new point shape files<br />

from the survey point dataset. In addition, kernel densities were<br />

generated for sightings in the Central region (Fig. 1) forwinter<br />

(as usually defined, above) and three other time periods: postwinter<br />

(March 16<strong>–</strong>April 30), summer (May 1<strong>–</strong>September 30),<br />

and pre-winter (October 1<strong>–</strong>November 14). The Central region<br />

is of particular importance because 1) its seagrass beds are the<br />

closest foraging grounds to the primary warm-water refuge at<br />

the FPL power plant (Fig. 2) and 2) most manatees that use<br />

this refuge must travel through the Central region. This region<br />

was analyzed separately and according to shorter time periods<br />

in order to identify fine-scale temporal and spatial changes in<br />

the distribution of manatees in this area.<br />

An output cell size of 90 and a search radius of 2500m 2<br />

were applied to both analyses to generate density estimates<br />

for each survey point. The distribution of data was best<br />

represented on maps by using quantile classifications of<br />

numeric data. This method creates classes that may be close<br />

together in terms of the values, but provides a scale that<br />

groups the majority of the data most appropriately given the<br />

density distribution of manatees at survey points. Color<br />

intensity was used to depict areas of high and low density.<br />

Management<br />

When large numbers or notable shifts in distribution of<br />

manatees occurred, counts and observations were reported to<br />

local and state managers and shared with law enforcement via<br />

the Lee County Marine Law Enforcement Task Force<br />

(LCMLETF). This group was formed in 2003 and includes<br />

members from the Lee County Sheriff’s Office, Fort Myers<br />

Police Department, Cape Coral Police Department, Sanibel<br />

Police Department, U.S. Coast Guard Station Fort Myers<br />

Beach, U.S. Coast Guard Cutter Marlin, USFWS Division of<br />

Law Enforcement, USFWS/J.N. “Ding” Darling National<br />

Wildlife Refuge, FWC Division of Law Enforcement Fort<br />

Myers Field Office, Florida Department of Environmental<br />

Protection Division of Law Enforcement, U. S. Power<br />

Squadrons, and U.S. Army Corps of Engineers. The group’s<br />

Mission Statement specifically addresses resource protection:<br />

“The agencies of the Lee County Marine Law<br />

Enforcement Task Force are committed to providing<br />

the highest quality of marine law enforcement to<br />

protect the users of Lee County’s waterways, safeguard<br />

property, and conserve/protect marine life<br />

along with its environment.” (Lee County Marine<br />

Law Enforcement Task Force 2004)<br />

This well-organized group meets monthly to coordinate<br />

and maximize efforts to enforce boating safety and marine<br />

wildlife regulations. The goal of providing survey information<br />

to this task force was to allow the leaders of state and<br />

local law enforcement to deploy officers in ways that<br />

optimized manatee protection, while also ensuring that<br />

other enforcement obligations were handled as well.<br />

Current speed zone regulations for the county were<br />

acquired (Fig. 3) and compared with the kernel density maps<br />

to evaluate the zonal boundaries, both year-round and<br />

seasonal, and the potential need for regulatory improvement.<br />

Results<br />

Survey counts<br />

A total of 34 flights with both the north and south surveys<br />

completed was considered for both spatial and numerical<br />

analysis; all incomplete flights (n=3) were removed from<br />

the dataset for this analysis. Each point (n=4,000)<br />

corresponded to a sighting of 1<strong>–</strong>102 manatees. The<br />

majority of sightings were single animals or small groups<br />

(mode=1, mean=3). Total counts among survey dates for<br />

northern and southern surveys combined ranged from 135


576 D. J. Semeyn et al.<br />

Fig. 1 Reference map of study<br />

area showing the locations of<br />

warm-water refuges at FPL<br />

Fort Myers, Matlacha Isles and<br />

Ten Mile Canal. The spatial<br />

extent of the Central region is<br />

indicated by the black rectangle.<br />

manatees to 626 manatees (Table 1). The mean number of<br />

manatees counted per flight was 345 (± 23). An independent<br />

T-test showed that mean counts were significantly<br />

different (P


Aerial surveys of manatees (Trichechus manatus) in Lee County, Florida 577<br />

Fig. 2 Map illustrating locations of seagrass beds throughout the entire<br />

study area (source: South Florida Water Management District 2003)<br />

(Fig. 5a, b). This movement coincides with a significant<br />

increase in mean number of manatees sighted per survey<br />

(p=0.00102). From pre-winter to winter, manatees<br />

became even more densely clustered (Fig. 5c), but the<br />

increase in the mean number of manatees per survey was<br />

not significantly different (p=0.582). During post-winter<br />

there were generally fewer high-density areas and fewer<br />

manatees sighted than there were for winter, indicating a<br />

dispersal of manatees as the weather warmed (Fig. 5d).<br />

Management<br />

With regard to our study, the LCMLETF provided a stable<br />

infrastructure for quick dissemination of individual flight<br />

results to enhance conservation efforts with the near real-time<br />

information gathered. The information we provided regarding<br />

distribution, density, and behavior of manatees was reported to<br />

resource managers and LCMLETF coordinators, and was<br />

subsequently used to position marine enforcement units in the<br />

most critical areas to provide additional protection in needed<br />

areas.<br />

A comparison between the Kernel density analyses<br />

(Figs. 4 and 5) and the county’s site-specific boat speed<br />

zones (Fig. 3) show that, with a few exceptions, areas with<br />

high densities of manatees year-round overlap well with<br />

year-round speed zones, and areas with high densities<br />

during “non-winter” months overlap remarkably well with<br />

the seasonally (Apr 1<strong>–</strong>Nov 15) designated zones. There<br />

appear to be relatively few areas where high densities of<br />

manatees are found year-round in seasonally designated<br />

zones and vice versa. While it would be in the interest of<br />

local and state officials to identify these areas where<br />

enforcement zones could be added or re-assessed, it is not<br />

within the scope of this paper to do so.<br />

Discussion<br />

Aerial survey counts<br />

The results reaffirm that Lee County supports large<br />

numbers of manatees, particularly during fall and winter<br />

months, and further emphasize the need for careful<br />

management of this particular population. These results<br />

can be compared to results from statewide synoptic surveys<br />

which also indicate that a large proportion of the state’s<br />

manatees is found in Lee County during winter. In the<br />

January 2010 synoptic survey, biologists counted 5,067<br />

manatees statewide, with 877 (approximately <strong>17</strong>%) in Lee<br />

County (H. Edwards FWC pers. comm.). This count was<br />

approximately 200 manatees more than the highest count in<br />

this study; however synoptic surveys are flown on days<br />

where sighting conditions are optimal. In February of the<br />

same year, after a prolonged period of cold weather, 905<br />

manatees were counted at the Orange River FPL plant alone<br />

(Reynolds, unpub. data).<br />

Our surveys produced much higher counts of manatees<br />

than did previous studies conducted in the area. The most<br />

recent study prior to ours was completed by FWC from<br />

1994 to 1998. For reference, the previous study reported an<br />

average of 147.6 manatees counted per survey (Fish and<br />

Wildlife Commission unpublished), whereas the mean<br />

count in our study was 345 manatees per survey. However,<br />

due to differences in flight path and overall survey effort,<br />

comparison between the 1994<strong>–</strong>1998 counts and ours is not<br />

straightforward. Regardless, both studies showed that<br />

manatees were present in Lee County in much higher<br />

numbers during winter than non-winter months. Also, the<br />

highest counts from both studies were recorded on surveys<br />

where large aggregations of manatees were present at<br />

warm-water sites, indicating that animals traveled to Lee<br />

County specifically for this resource and/or that manatees<br />

aggregated at the plant are easier to count than smaller<br />

groups of widely dispersed manatees. Survey data from the<br />

Central region also showed that there were more manatees<br />

present during winter than non-winter. However, over


578 D. J. Semeyn et al.<br />

Fig. 3 Reference map of major<br />

speed zones within Lee County<br />

(source: Florida Fish and<br />

Wildlife Conservation<br />

Commission- Fish and Wildlife<br />

Research Institute)<br />

shorter time periods, analyses support the observation that<br />

the number of manatees in this region during pre-winter<br />

was not significantly different from winter numbers. The<br />

aggregation of manatees at a warm-water refuge or nearby<br />

seagrass beds provides an opportunity for protection of<br />

manatees during pre-winter and winter that is more difficult<br />

to achieve when the animals are dispersed during nonwinter<br />

months.<br />

Spatial analysis and management implications<br />

The figures depicting manatee densities provide a clear<br />

comparison of spatial and temporal changes in numbers and<br />

habitat use in Lee County. This information has important<br />

management implications. For example, two key months,<br />

October and November, mark the influx of tourists and<br />

seasonal residents to southwestern Florida. Density analyses<br />

showed that during these months, the distribution of manatees<br />

was highly concentrated within San Carlos Bay and Matlacha<br />

Pass.Thispre-winterdistributionismorelikethatofthe<br />

winter than the summer, suggesting that manatees may be<br />

positioning or “staging” themselves for quick access to<br />

essential warm-water habitats well before cold weather<br />

arrives. Bengtson (1981) noted a similar strategy where<br />

manatees moved closer to Blue Spring, Florida in the fall to<br />

be nearer to a warm water refuge. These large concentrations,<br />

however, may indicate more than just positioning<br />

for imminent movement to warm water. Given the fact that<br />

San Carlos Bay contains abundant seagrass beds, it is<br />

possible that manatees use this pre-winter time to feed in<br />

preparation for bouts of fasting during potentially prolonged<br />

residential periods within warm-water refuge (Rommel et al.<br />

2003). In any case, the coincidental arrival of expanded<br />

numbers of both people and manatees to specific waterways<br />

presents a significant challenge for enforcement, and<br />

limitations in the size of enforcement staff underscore the


Aerial surveys of manatees (Trichechus manatus) in Lee County, Florida 579<br />

Table 1 Total number of manatees observed on each of 34 flights,<br />

and total number of manatees observed in the Central region<br />

Survey date Survey total Central region total<br />

4/2/2007 158 24<br />

4/24/2007 <strong>17</strong>9 44<br />

5/<strong>17</strong>/2007 216 70<br />

5/29/2007 235 52<br />

6/8/2007 203 18<br />

6/28/2007 207 56<br />

7/2/2007 216 78<br />

7/9/2007 220 96<br />

8/13/2007 242 90<br />

8/20/2007 135 37<br />

9/7/2007 234 84<br />

10/8/2007 203 90<br />

11/6/2007 398 180<br />

11/20/2007 499 211<br />

12/3/2007 504 233<br />

12/18/2007 396 163<br />

1/7/2008 416 89<br />

2/19/2008 465 193<br />

4/<strong>17</strong>/2008 353 1<strong>17</strong><br />

5/13/2008 414 104<br />

5/27/2008 349 105<br />

6/6/2008 357 125<br />

7/18/2008 244 106<br />

8/28/2008 288 96<br />

10/8/2008 357 187<br />

11/5/2008 497 294<br />

12/8/2008 420 151<br />

12/18/2008 485 283<br />

1/9/2009 518 188<br />

2/27/2009 626 328<br />

3/9/2009 610 211<br />

3/30/2009 411 127<br />

4/10/2009 261 91<br />

4/29/2009 397 113<br />

Total 1<strong>17</strong>13 4434<br />

Table 2 Lee County survey study area: total number of surveys,<br />

manatees sighted, and the mean number of manatees sighted ± one<br />

standard error per survey for each specified time period<br />

Winter Non-winter<br />

# of Surveys 11 25<br />

Total # Manatees 4939 6774<br />

Mean #/Survey 494±24 282±20<br />

Table 3 Central Region: total number of surveys, manatees sighted ± one<br />

standard error, and the mean number of manatees sighted per survey for<br />

each specified time period<br />

Pre-winter Winter Post-winter Summer<br />

# of Surveys 4 10 6 14<br />

Total # of Manatees 751 2050 516 2384<br />

Mean #/Survey 188±42 205±21 86±<strong>17</strong> 80±8<br />

need to operate as efficiently as possible. Therefore, the<br />

identification of important seasonal habitats where management<br />

and enforcement efforts should be focused is an<br />

essential outcome of this study.<br />

Spatial data for the Central region also show an<br />

interesting post-winter pattern. At this time, manatees focus<br />

on feeding, but their distributions are more similar to<br />

summer patterns, which are more widely distributed. When<br />

cold weather ceases, manatees appear to disperse quickly,<br />

returning to forage in areas other than those within San<br />

Carlos Bay and Matlacha Pass. This rapid dispersal pattern<br />

has previously been noted as a contributing factor to the<br />

mass mortality of manatees and other marine animals in<br />

southwest Florida during a 1996 red-tide event (Landsberg<br />

and Steidinger 1998). This paper concluded that manatees are<br />

at high risk from February through April should a “perfect<br />

storm” of environmental factors combine to generate persistent<br />

red-tide blooms in the region. Exposure to the neurotoxins<br />

produced by the bloom-forming dinoflagellate, Karenia<br />

brevis, causes disorientation and severe health problems in<br />

manatees (O’Shea et al. 1991; Bossart et al. 1998), making<br />

them particularly vulnerable to boat strikes. While little can<br />

apparently be done to prevent manatees from encountering<br />

these blooms, increased protection against human-related<br />

threats during such events should be a priority for managers.<br />

Manatee travel patterns within seasons are also of concern<br />

to resource managers. For example, surveyors observed that<br />

on the coldest winter days most manatees were found at<br />

warm-water sites, whereas on warmer days between cold<br />

fronts many of them were found in San Carlos Bay,<br />

presumably foraging. In order to make this transition from<br />

one critical resource to another, manatees must travel a<br />

minimum of 38 km up or down the Caloosahatchee River.<br />

This is a concern for managers since the primary travel<br />

corridor for manatees coincides with the primary channel for<br />

boat traffic. Even more troublesome, a 1998 study assessing<br />

the effectiveness of speed zones at multiple sites within this<br />

river yielded an overall compliance of just 58% (Gorzelany<br />

2004). Although this study considered only a subset of Lee<br />

County waters, without evidence to the contrary it can be<br />

assumed that compliance throughout the rest of the county is<br />

similar to that of the Caloosahatchee. On a more positive<br />

note, Gorzelany (2004) also reported a significant increase in


580 D. J. Semeyn et al.<br />

Fig. 4 Kernel density analysis of distribution of manatee sightings within the entire study area for (a) non-winter (16 March<strong>–</strong>14 November) and<br />

(b) winter (15 November<strong>–</strong>15 March)<br />

compliance levels when law enforcement was present. The<br />

results from his boater compliance study imply that our<br />

communication efforts with the LCMLETF not only efficiently<br />

directed limited enforcement staff to areas of potential<br />

watercraft incidents, but also indirectly increased the<br />

likelihood of boaters obeying the posted regulations.<br />

Management in action<br />

Watercraft-related mortality within Lee County did not<br />

decrease over the course of this study. In their evaluation of<br />

the impact of boat speed restrictions on watercraft-related<br />

deaths, Laist and Shaw (2006) listed four reasons for a lack<br />

of decline in deaths state-wide after implementation of speed<br />

zones in 13 key counties: (1) manatees are unable to avoid<br />

slow-moving vessels, and therefore speed restrictions offer<br />

little protection (2) low boater compliance occurs within<br />

speed zones (3) current speed zones are not substantial<br />

enough to offer protection (4) increasing numbers of both<br />

boaters and manatees result in increase numbers of collisions,<br />

effectively outpacing the speed zone reductions of<br />

these collisions. The Laist and Shaw study revealed that<br />

implementing slow speed rules within two densely populated<br />

manatee areas substantially decreased watercraft-related<br />

deaths. Based on their findings and those from our study,<br />

which indicated that the federal and state speed zones are<br />

spatially and seasonally substantiated in Lee County, the first<br />

and third reasons for mortality not decreasing seem unlikely.<br />

We conclude that in the case of Lee County, the absence of a<br />

decline in watercraft-related mortality is most likely explained<br />

by an increase in numbers of both manatees and boaters and/or,<br />

as discussed above, low boater compliance.<br />

One of the major benefits realized from the original<br />

objective of the study was the incorporation of new spatial<br />

data into the evaluation process for proposed marine facilities.<br />

The new spatial data from our study provide managers with<br />

up-to-date information reflecting the distribution of manatees<br />

in Lee County, which is important for the siting of marine<br />

facilities. In addition, prompt, near real-time reporting resulted<br />

in excellent communication and partnership among biologists,<br />

managers and law enforcement. These efforts helped to<br />

enhance both awareness of manatee presence and protection<br />

needs and the enforcement of posted manatee zones, which<br />

will hopefully result in reduced mortality rates in the<br />

upcoming years. Ultimately, the law enforcement officers are<br />

among the most important end users of the results of our<br />

research, as they have more daily contact with the boating<br />

public than biologists or decision makers. In fiscal year 2008,<br />

four Lee County marine law enforcement agencies conducted<br />

2,665 non-citation or educational contacts (Lee County<br />

Department of Natural Resources, unpublished data). This<br />

demonstrates the tremendous potential educational impact<br />

that law enforcement has on the boating public. To take<br />

advantage of this opportunity, it is the responsibility of


Aerial surveys of manatees (Trichechus manatus) in Lee County, Florida 581<br />

Fig. 5 Kernel density analysis of the distribution of manatee sightings within the Central region for four time periods: (a) summer (01 May<strong>–</strong>30<br />

September); (b) pre-winter (01 October<strong>–</strong>14 November); (c) winter (15 November<strong>–</strong>15 March); and (d) post-winter (16 March<strong>–</strong>30 April)


582 D. J. Semeyn et al.<br />

biologists and other researchers to provide law enforcement<br />

with the most timely, accurate, and understandable<br />

information possible.<br />

Open dialogue among biologists, resource managers, and<br />

law enforcement officers is critical to reaching a balance<br />

between ensuring human access to waterways and resources<br />

contained therein and providing protection for vulnerable<br />

natural resources. This coordination is especially important in<br />

areas such as Lee County where there is high use of key<br />

habitats by both people and manatees, and where tension<br />

exists surrounding regulations aimed at protecting a species or<br />

habitat. Open dialogue fosters partnerships, mutual respect,<br />

and an understanding that is crucial for manatee conservation<br />

and protection. Lee County and the LCMLETF is a model for<br />

this type of partnership.<br />

Conclusions<br />

With minimal cost and time, this project contributed to<br />

conservation and management of manatees in Lee County.<br />

It provided:<br />

□ aerial survey data that clarified both the larger-thanexpected<br />

numbers of manatees present in Lee County<br />

waters and the habitats they prefer seasonally;<br />

□ a spatial and temporal monitoring tool for long term<br />

decision making with regard to balancing the development<br />

of coastal waterways and protection of key areas<br />

for manatees;<br />

□ real-time information on manatee numbers and locations<br />

to enhance enforcement actions;<br />

□ validation of current state regulatory zones set within<br />

the county.<br />

Possibly the largest impact was in the simplest aspect of<br />

the project: communication. Through building partnerships<br />

and interagency cooperation, studies such as this one can<br />

hope to promote conservation at a much higher level. Not<br />

only did it provide managers with information to make<br />

sound decisions when considering regulations or permits in<br />

a dynamic environment, but it allowed law enforcement to<br />

maximize limited resources and focus on the most critical<br />

areas.<br />

Resource managers benefit from the knowledge of<br />

when and where to be most vigilant when navigating the<br />

waterways of Lee County. This information should be<br />

relayed to the public through a variety of media outlets<br />

in several different mediums. Future research in any of<br />

the 13 key counties should examine distribution of<br />

manatees compared to existing regulatory zones and<br />

focus on communication between agencies and with the<br />

public to further conservation efforts of this endangered<br />

species.<br />

Acknowledgements We thank the pilots and observers that conducted<br />

the surveys including Lew Lawrence, Dr. James Powell, Greg Baker,<br />

Jorge Neumann, and Kat Frisch. We are especially grateful to the staff of<br />

Florida Power & Light Company, particularly Winifred Perkins and Jodie<br />

Gless, for their patience and support throughout the course of the study as<br />

they coordinated our surveys with FPL security staff. Special appreciation<br />

also goes to the air traffic controllers at Fort Myers Page Field and<br />

Southwest Florida <strong>International</strong> Airport. Thanks also go to the Florida<br />

Fish and Wildlife Conservation Commission Law Enforcement officers<br />

and the Lee County Marine Law Enforcement Task Force for their<br />

interest and cooperation. The authors would also like to thank Jay<br />

Sprinkel for his advice and review of the statistical analysis, Kimberly<br />

Miller, Alexis Levengood, and Adrien Miller for their help editing and<br />

formatting the manuscript, and Holly Edwards for providing information<br />

on present and previous studies conducted by the FWC in Lee County.<br />

Funding for this project was provided by the Lee County Department of<br />

Natural Resources, and airplane fuel was kindly donated by Dolphin<br />

Aviation of Sarasota, Florida.<br />

References<br />

Ackerman BB, Wright SD, Bonde RK, O’Dell DK, Banowetz DJ<br />

(1995) Trends and patterns in mortality of manatees in Florida,<br />

1974<strong>–</strong>1992. In: O’Shea TJ, Ackerman BB, Percival HF (eds)<br />

Population biology of the Florida manatee. U.S. Department of<br />

the Interior, National Biological Service, Information and<br />

Technology Report 1, pp 223<strong>–</strong>258<br />

Bengtson JL (1981) Ecology of manatees (Trichechus manatus) in the St.<br />

John’s River,Florida.Master’s Thesis, University of Minnesota<br />

Bossart GD, Baden DG, Ewing RY, Roberts B, Wright SD (1998)<br />

Brevetoxicosis in manatees (Trichechus manatus latirostris) from<br />

the 1996 epizootic: gross, histological, and immunohistochemical<br />

features. Environ Toxicol Pathol 26:276<strong>–</strong>282<br />

Craig BA, Reynolds JE III (2004) Determination of manatee<br />

population trends along the Atlantic coast of Florida using a<br />

Bayesian approach with temperature-adjusted aerial survey data.<br />

Mar Mamm Sci 20:386<strong>–</strong>400<br />

Florida Fish and Wildlife Conservation Commission, Division of Law<br />

Enforcement (2008) Boating Accidents Statistical Report pp 23<br />

http://www.myfwc.com/docs/Safety/2008_Boating_Statbook.pdf.<br />

Accessed December 2009<br />

Florida Fish and Wildlife Conservation Commission, Fish and<br />

Wildlife Research Institute (2010) Manatee Mortality Database<br />

1974<strong>–</strong>2009. http://research.myfwc.com/features/category_sub.<br />

asp?id=2241. Accessed January 2010<br />

Gorzelany JF (2004) Evaluation of boater compliance with manatee speed<br />

zones along the Gulf coast of Florida. Coast Manage 32:215<strong>–</strong>226<br />

Horne JS, Garton EO (2006) Likelihood cross-validation versus least<br />

squares cross-validation for choosing the smoothing parameter in<br />

kernel home-range analysis. J Wildl Manage 70:641<strong>–</strong>648<br />

Laist DW, Shaw C (2006) Preliminary evidence that boat speed restrictions<br />

reduce deaths of Florida manatees. Mar Mamm Sci 22:472<strong>–</strong>479<br />

Landsberg JH, Steidinger KA (1998) A historical review of Gymnodinium<br />

breve red tides implicated in mass mortalities of the manatee<br />

(Trichechus manatus latirostris) in Florida, USA. In: Reguera B,<br />

Blanco J, Fernández ML, Wyatt T (eds) Harmful algae. Xunta de<br />

Galicia and IOC of UNESCO Publishers, p 97<strong>–</strong>100<br />

Langtimm CA, Beck CA (2003) Lower survival probabilities for adult<br />

Florida manatees in years with intense coastal storms. Ecol Appl<br />

13:257<strong>–</strong>268<br />

Langtimm CA, Beck CA, Edwards HH, Fick-Child KJ, Ackerman<br />

BB, Barton SL, Hartley WC (2004) Survival estimates for<br />

Florida manatees from the photo-identification of individuals.<br />

Mar Mamm Sci 20:438<strong>–</strong>463


Aerial surveys of manatees (Trichechus manatus) in Lee County, Florida 583<br />

Lee County Marine Law Enforcement Task Force (2004) http://www.<br />

marinetaskforce.com. Accessed January 2010<br />

O’Shea TJ, Rathbun GB, Bonde RK, Buergelt CD, Odell DK (1991)<br />

An epizootic of Florida manatees associated with a dinoflagellate<br />

bloom. Mar Mamm Sci 7:165<strong>–</strong><strong>17</strong>9<br />

Reynolds JE III, Morales-Vela B, Lawler I, Edwards H (2010) Utility<br />

and design of aerial surveys for sirenians. In: Hines E, Reynolds<br />

JE III, Mignucci-Giannoni AA, Aragones LV, Marmontel M<br />

(eds) <strong>Sirenian</strong> conservation: issues and strategies in developing<br />

countries. University Press of Florida, Gainesville<br />

Rommel SA, Reynolds JE III, Lynch HA (2003) Adaptations of the<br />

herbivorous marine mammals In: Mannetje L, Ramírez-Avilás L,<br />

Sandoval-Castro C, Ku-Vera JC (eds) Proceedings of an <strong>International</strong><br />

Symposium on the Nutrition of Herbivores pp 287<strong>–</strong>306<br />

Seaman DE, Powell RA (1996) An evaluation of the accuracy of<br />

kernel density estimators for home range analysis. Ecology<br />

77:2075<strong>–</strong>2085<br />

United States Census Bureau (2009) Lee County Florida. http://quickfacts.<br />

census.gov/qfd/states/12/12071.html. Accessed November 2009<br />

United States Fish and Wildlife Service (2009) Department of the<br />

Interior. Marine Mammal Protection Act, Stock Assessment<br />

Report. http://www.fws.gov/northflorida/Manatee/SARS/<br />

20091230_frn_NOA_Manatee_Final_SARs_74_fr_69136.pdf.<br />

Accessed January 2010<br />

USA Cities Online (2009) Lee County Florida. http://www.<br />

usacitiesonline.com/flleecounty.htm Accessed December 2009<br />

Worton BJ (1995) Using Monte Carlo simulation to evaluate<br />

kernel-based home range estimators. J Wildl Manage 59:794<strong>–</strong><br />

800<br />

Wright IE, Reynolds JE III, Ackerman BB, Ward LI, Weigle BL,<br />

Szelistowski WA (2002) Trends in manatee (Trichechus<br />

manatus latirostris) counts and habitat use in Tampa Bay,<br />

1987<strong>–</strong>1994: implications for conservation. Mar Mamm Sci<br />

18:259<strong>–</strong>274


Intraspecific and geographic variation of West Indian manatee<br />

(Trichechus manatus spp.) vocalizations (L)<br />

Douglas P. Nowacek<br />

Biology Department, Woods Hole Oceanographic Institution, Woods Hole, Massachusetts 02543 and<br />

Sensory Biology and Behavior Program, Mote Marine Laboratory, 1600 Ken Thompson Parkway, Sarasota,<br />

Florida 34236<br />

Brandon M. Casper<br />

College of Marine Science, University of South Florida, 140 Seventh Avenue South, St. Petersburg,<br />

Florida 33701<br />

Randall S. Wells and Stephanie M. Nowacek<br />

Chicago Zoological Society, c/o Mote Marine Laboratory, 1600 Ken Thompson Parkway, Sarasota,<br />

Florida 34236<br />

David A. Mann a)<br />

College of Marine Science, University of South Florida, 140 Seventh Avenue South, St. Petersburg,<br />

Florida 33701 and Sensory Biology and Behavior Program, Mote Marine Laboratory,<br />

1600 Ken Thompson Parkway, Sarasota, Florida 34236<br />

Received 25 October 2003; revised 11 April 2003; accepted 14 April 2003<br />

Recordings of manatee Trichechus manatus spp. vocalizations were made in Florida and Belize to<br />

quantify both intraspecific and geographic variation. Manatee vocalizations were relatively<br />

stereotypical in that they were short tonal harmonic complexes with small frequency modulations at<br />

the beginning and end. Vocalizations ranged from almost pure tones to broader-band tones that had<br />

a raspy quality. The loudest frequency was typically the second or third harmonic, with average<br />

received levels of the peak frequency of about 100 dB re 1 Pa. Signal parameters measured from<br />

these calls showed the manatees from Belize and Florida have overlapping distributions of sound<br />

duration, peak frequency, harmonic spacing, and signal intensity, indicating no obvious<br />

distinguishing characteristics between these isolated populations. © 2003 Acoustical Society of<br />

America. DOI: 10.1121/1.1582862<br />

PACS numbers: 43.80.Ka, 43.30.Sf, 43.80.Ev WA<br />

I. INTRODUCTION<br />

Florida manatees, Trichechus manatus latirostris, are endangered,<br />

and many animals are killed or injured each year<br />

by boat strikes. At least 25% of annual documented manatee<br />

deaths are due to collisions with vessels Marine Mammal<br />

Commission, 2002. Despite the fact that behavioral evidence<br />

indicates that manatees do have the ability to detect<br />

and respond to approaching vessels Nowacek et al., 2000;<br />

Weigle et al., 1994, the animals continue to be hit. In an<br />

effort to contribute to solutions that might reduce manatee<br />

morbidity and mortality from boat strikes, we set out to build<br />

a device to alert boaters of the presence of manatees based<br />

on passive detection of their sounds.<br />

Most of the boat strike mitigation effort has been directed<br />

toward testing the hearing capabilities of manatees<br />

Gerstein et al., 1999; Ketten et al., 1992 and their behavioral<br />

response in the presence of boats Nowacek et al.,<br />

2000; Weigle et al., 1994 in hopes of understanding if and<br />

how they respond to oncoming vessels. Manatee sound production<br />

has been documented in only four papers Evans and<br />

Herald, 1970; Schevill and Watkins, 1965; Sonoda and Take-<br />

a<br />

Author to whom correspondence should be addressed. Electronic mail:<br />

dmann@marine.usf.edu<br />

mura, 1973; Sousa-Lima et al., 2002, including very little<br />

discussion of variability inherent in the animals’ vocal repertoire.<br />

These reports describe several types of sounds, but the<br />

primary vocalization recorded was a tonal sound often having<br />

several harmonics with the second or third harmonic often<br />

stronger than the fundamental frequency; intraspecific<br />

variation was not reported. The fundamental frequency<br />

ranged from 2.5 to 5 kHz for the West Indian manatee,<br />

Trichechus manatus spp., and 2.6 to 5.9 kHz for the Amazonian<br />

manatee, Trichechus inunguis.<br />

Vocalizations may be the most easily detected sounds<br />

produced by manatees compared to chewing and flatulence<br />

because published reports show them to have the highest<br />

signal-to-noise ratio and the vocalizations occur in frequency<br />

bands most dissimilar to the natural background noise of the<br />

manatees’ environment. To determine the amount of stereotypy<br />

in vocalizations we recorded sounds from Antillean<br />

manatees Trichechus manatus manatus in Southern Lagoon,<br />

Belize Lat/Lon: <strong>17</strong>° 12N, 88° 20W) and Florida<br />

manatees in Crystal River, FL Lat/Lon: 28° 53 N, 82°<br />

35W). These two groups of manatees are subspecies of the<br />

West Indian manatee, so comparing the structure of their<br />

vocalizations could add to our understanding of the similarities<br />

and differences between the two subspecies.<br />

66 J. Acoust. Soc. Am. 114 (1), July 2003 0001-4966/2003/114(1)/66/4/$19.00 © 2003 Acoustical Society of America


II. METHODS<br />

A. Belize recordings<br />

In Belize we recorded sounds on a digital archival tag,<br />

the DTAG Johnson and Tyack, 2003, which simultaneously<br />

records the attitude pitch and roll, heading, and depth of the<br />

animal with sounds produced or received by the animal. The<br />

hydrophone signal was digitized continuously at 32 kHz. Selected<br />

manatees were encircled by a seine net in shallow<br />

water and brought aboard the capture vessel for health assessment,<br />

measurements, and tagging. DTAGs were attached<br />

to three individual manatees via temporary attachments using<br />

standard manatee peduncle belts. Floats tethered to the belts<br />

carried VHF transmitters for direct radio-tracking, and in<br />

some cases a GPS receiver. Sonic tags incorporated into the<br />

FIG. 2. Power spectrum of a typical manatee vocalization. Plot shows a<br />

2048-point FFT of a 2048-point segment from the middle of the signal<br />

shown in Fig. 1a. * indicates the fundamental frequency.<br />

belts permitted underwater tracking. The tags were recovered<br />

either when the belts released via a corrodible link or animals<br />

were recaptured for recovery of the DTAG and removal of<br />

other tags and belts. Attachments lasted approximately 24 h<br />

with acoustic data recorded for 4 or 8 h depending on the tag<br />

settings used.<br />

B. Crystal River, FL recordings<br />

Recordings in Crystal River, FL were made using a hydrophone<br />

HTI 96-min; sensitivity 164 dBV/Pa, 20 Hz to<br />

32 kHz connected to a data acquisition system Tucker-<br />

Davis Technologies, RP2.1 24-bit ADC and laptop computer.<br />

Signals were acquired continuously for approximately<br />

2 h at 48 828.125 Hz sample rate from an anchored boat with<br />

the hydrophone on the bottom at 1-m depth. Recordings<br />

were made approximately 20 m away from a group of about<br />

50 manatees that were in the spring. Individuals would occasionally<br />

move out of the spring into the spring run to<br />

within 2moftheboat.<br />

C. Data analysis<br />

FIG. 1. Spectrograms of representative<br />

manatee vocalizations. a Typical<br />

tonal harmonic vocalization Crystal<br />

River, FL, b tonal vocalization transitioning<br />

to less tonal Crystal River,<br />

FL, c broader-band harmonic vocalization<br />

Crystal River, FL, and d<br />

typical tonal harmonic vocalization<br />

Southern Lagoon, Belize. The acoustic<br />

tag used to record manatees in Belize<br />

sampled at F’s32 kHz. The<br />

scale bar shows relative sound levels<br />

in decibels.<br />

Data were analyzed with MATLAB Mathworks, Inc..<br />

Peak frequency frequency with the most energy and the<br />

level of the peak frequency were determined by performing a<br />

1024-point FFT on a 1024-point segment of the middle of<br />

each call. This minimized smearing of frequencies due to<br />

changes in frequencies during the call. Harmonic spacing<br />

which is equivalent to the fundamental frequency was determined<br />

by performing an autocorrelation of the FFT, and<br />

measuring the first peak after lag zero. Duration was measured<br />

by MATLAB with an automatic detection algorithm,<br />

and then verified by hand from the spectrogram. While this<br />

may lead to a small uncertainty in measuring the time, it was<br />

advantageous for signals with low signal-to-noise ratios.<br />

J. Acoust. Soc. Am., Vol. 114, No. 1, July 2003 Nowacek et al.: Letters to the Editor<br />

67


III. RESULTS<br />

Manatees in Florida and Belize produced vocalizations<br />

that were harmonic complexes with small frequency modulations<br />

at the beginning and end Fig. 1. These ranged from<br />

almost pure tones to broader-band tones that have a noisy<br />

quality. The loudest frequency was typically the second or<br />

FIG. 3. Distribution of manatee vocalization<br />

parameters from a Crystal<br />

River, FL and b Southern Lagoon,<br />

Belize data from 11 March 2002.<br />

Panels show peak frequency, level of<br />

peak frequency, fundamental frequency<br />

harmonic spacing, and duration<br />

from top to bottom. We show only<br />

the Belize data from 11 March because<br />

this animal was in a similar behavioral<br />

situation to those in Crystal River. The<br />

animal was with a group of manatees,<br />

so these samples include sounds from<br />

other manatees as well as from the<br />

tagged animal. We felt that comparing<br />

sounds produced in similar situations<br />

provided the most accurate comparison<br />

for our study.<br />

third harmonic Figs. 2 and 3. Signal parameters measured<br />

from these calls show the manatees from Belize and Florida<br />

have overlapping distributions of sound duration, peak frequency,<br />

harmonic spacing, and signal intensity Table I and<br />

Fig. 3. Statistical comparisons between these signals were<br />

not possible because it is not known which individuals pro-<br />

68 J. Acoust. Soc. Am., Vol. 114, No. 1, July 2003 Nowacek et al.: Letters to the Editor


duced which sounds. Treating each call as a separate replicate<br />

would lead to pseudoreplication. The rate of vocalization<br />

in Crystal River, accounting for the number of animals<br />

present, was 1.29 vocalizations per minute, and in Belize it<br />

ranged from 0.09 to 0.75 per minute for the three animals<br />

tagged. When alone, the Belize animals were often silent for<br />

periods of 10 min. The rates of vocalization we measured<br />

were similar to earlier reports Bengtson and Fitzgerald,<br />

1985.<br />

IV. DISCUSSION<br />

TABLE I. Basic statistics of sounds recorded from manatees in Southern Lagoon, Belize and Crystal River, FL.<br />

The mean for each data set is shown with standard deviation in parentheses.<br />

Belize<br />

9 March<br />

2002<br />

There are no obvious differences in the vocalizations<br />

produced by manatees from Florida and Belize. For the parameters<br />

that were characterized sound duration, peak frequency,<br />

and harmonic spacing manatees from Florida and<br />

Belize had overlapping distributions. It is possible, however,<br />

that there are differences that we did not characterize.<br />

The finding that the second or third harmonic of the<br />

vocalization is usually most intense could be due either to<br />

how the sound is produced, or to propagation effects where<br />

the lower frequencies do not propagate as well in shallow<br />

water, or a combination of the two Rogers and Cox, 1988.<br />

Given that previous research on captive animals also found<br />

the fundamental to be less intense Evans and Herald, 1970;<br />

Schevill and Watkins, 1965, our results are probably best<br />

explained by the production system of the animals. Indeed,<br />

in all of these recordings we do not know the distance to the<br />

sound-producing manatee, only a range to the manatees that<br />

were in the area. The data from Belize include both the animal<br />

wearing the DTAG as well as animals vocalizing nearby.<br />

Still the received levels can be taken to show the range of<br />

levels that might be produced by manatees, and they are<br />

likely within 6<strong>–</strong>15 dB of source levels given the range of<br />

distances over which the recordings were made assuming<br />

losses of 3 dB per doubling of distance.<br />

The motivation for this study was to determine the range<br />

of natural variability in manatee sounds, so that a passive<br />

acoustic detection device can be developed to warn boaters<br />

of the presence of manatees. These data show that West Indian<br />

manatee sounds are relatively stereotypical, even be-<br />

Belize<br />

10 March<br />

2002<br />

Belize<br />

11 March<br />

2002<br />

Crystal<br />

River<br />

n 26 105 208 218<br />

Peak frequency Hz 3180 727 7080 2207 5560 2559 5223 1937<br />

Peak level<br />

dB re 1 Pa<br />

97.1 4.3 92.5 6.6 100.0 4.7 103.6 6.8<br />

Fundamental<br />

frequency Hz<br />

3180 728 4380 1618 3630 1620 2867 1059<br />

Duration s 0.032 0.0<strong>17</strong> 0.161 0.10 0.2<strong>17</strong> 0.098 0.228 0.074<br />

tween subspecies, in that they are short tonal harmonic complexes,<br />

and lend themselves readily to this application.<br />

ACKNOWLEDGMENTS<br />

We thank John E. Reynolds III, Buddy Powell, Robert<br />

Bonde, Mesha Gough, and Kevin Andrewyn for their assistance<br />

in the field and/or their consultation on this manuscript<br />

and Mark Johnson and Alex Shorter for their assistance preparing<br />

and deploying the DTAGs. This project was supported<br />

by the Florida Fish and Wildlife Conservation Commission,<br />

Wildlife Trust, the Chicago Board of Trade, and the Chicago<br />

Zoological Society.<br />

Bengtson, J. L., and Fitzgerald, S. M. 1985. ‘‘Potential role of vocalizations<br />

in West Indian manatees,’’ J. Mammal. 664, 816<strong>–</strong>819.<br />

Evans, W. E., and Herald, E. S. 1970. ‘‘Underwater calls of a captive<br />

Amazon manatee, Trichechus inunguis,’’ J. Mammal. 51, 820<strong>–</strong>823.<br />

Gerstein, E. R., Gerstein, L., Forsythe, S. E., and Blue, J. E. 1999. ‘‘The<br />

underwater audiogram of the West Indian manatee Trichechus manatus,’’<br />

J. Acoust. Soc. Am. 105, 3575<strong>–</strong>3583.<br />

Johnson, M. P., and Tyack, P. L. 2003. ‘‘A digital acoustic recording tag for<br />

measuring the response of wild marine mammals to sound,’’ IEEE J.<br />

Ocean. Eng. 281, 3<strong>–</strong>12.<br />

Ketten, D. R., Odell, D. K., and Domning, D. P. 1992. Marine Mammal<br />

Sensory Systems, edited by J. Thomas Plenum, New York, pp. 77<strong>–</strong>95.<br />

Nowacek, S. M., Wells, R. S., Nowacek, D. P., Owen, E. C. G., Speakman,<br />

T. R., and Flamm, R. O. 2000. ‘‘Manatee behavioral responses to vessel<br />

approaches,’’ Report No. FWC-00127, 2000.<br />

Rogers, P. H., and Cox, M. 1988. Sensory Biology of Aquatic Animals,<br />

edited by J. Atema, A. N. Popper, and R. R. Fay Springer-Verlag, New<br />

York.<br />

Schevill, W. E., and Watkins, W. A. 1965. ‘‘Underwater calls of Trichechus<br />

Manatee,’’ Nature London 2054969, 373<strong>–</strong>374.<br />

Sonoda, S., and Takemura, A. 1973. ‘‘Underwater sounds of the manatees,<br />

Trichechus manatus and T. inunguis Trichechidae,’’ Report of the Institute<br />

for Breeding Research, Tokyo University of Agriculture, Vol. 4, pp.<br />

19<strong>–</strong>24.<br />

Sousa-Lima, R. S., Paglia, A. P., and Da Fonseca, G. A. B. 2002. ‘‘Signature<br />

information and individual recognition in the isolation calls of Amazonian<br />

manatees, Trichechus inunguis Mammalia: Sirenia,’’ Anim. Behav.<br />

63, 301<strong>–</strong>310.<br />

U. S. Marine Mammal Commission 2002. ‘‘Annual Report of the U.S.<br />

Marine Mammal Commission,’’ U.S. Marine Mammal Commission, 4340<br />

East West Highway, Suite 905, Bethesda, MD.<br />

Weigle, B. L., Wright, I. E., and Huff, J. A. 1994. ‘‘Responses of manatees<br />

to an approaching boat: a pilot study,’’ presented at the First international<br />

manatee and dugong research conference, Gainesville, FL.<br />

J. Acoust. Soc. Am., Vol. 114, No. 1, July 2003 Nowacek et al.: Letters to the Editor<br />

69


Genetica (2011) 139:833<strong>–</strong>842<br />

DOI 10.1007/s10709-011-9583-z<br />

Evidence of two genetic clusters of manatees with low genetic<br />

diversity in Mexico and implications for their conservation<br />

Coralie Nourisson • Benjamín Morales-Vela • Janneth Padilla-Saldívar •<br />

Kimberly Pause Tucker • AnnMarie Clark • Leon David Olivera-Gómez •<br />

Robert Bonde • Peter McGuire<br />

Received: 25 February 2011 / Accepted: 18 May 2011 / Published online: <strong>17</strong> June 2011<br />

Ó Springer Science+Business Media B.V. 2011<br />

Abstract The Antillean manatee (Trichechus manatus<br />

manatus) occupies the tropical coastal waters of the<br />

Greater Antilles and Caribbean, extending from Mexico<br />

along Central and South America to Brazil. Historically,<br />

manatees were abundant in Mexico, but hunting during the<br />

pre-Columbian period, the Spanish colonization and<br />

throughout the history of Mexico, has resulted in the significantly<br />

reduced population occupying Mexico today.<br />

The genetic structure, using microsatellites, shows the<br />

presence of two populations in Mexico: the Gulf of Mexico<br />

(GMx) and Chetumal Bay (ChB) on the Caribbean coast,<br />

with a zone of admixture in between. Both populations<br />

show low genetic diversity (GMx: NA = 2.69; HE = 0.41<br />

C. Nourisson (&) B. Morales-Vela J. Padilla-Saldívar<br />

El Colegio de la Frontera Sur, Av. Centenario Km 5.5,<br />

77000 Chetumal, Quintana Roo, Mexico<br />

e-mail: coralie.nourisson@gmail.com<br />

K. P. Tucker<br />

Department of Natural Sciences, College of Coastal Georgia,<br />

3700 Altama Avenue, Brunswick, GA 31520, USA<br />

A. Clark<br />

ICBR Genetic Analysis Laboratory, University of Florida,<br />

2033 Mowry Road, Gainesville, FL 32610, USA<br />

L. D. Olivera-Gómez<br />

Division Academica de Ciencias Biologicas, Universidad Juarez<br />

Autonoma de Tabasco, Km. 0.5 carretera Villahermosa-<br />

Cárdenas, 86039 Villahermosa, Tabasco, Mexico<br />

R. Bonde<br />

US Geological Survey, 2201 NW 40th Terrace, Gainesville,<br />

FL 32605, USA<br />

P. McGuire<br />

University of Florida, 2015 SW 16th Ave, Gainesville,<br />

FL 32610, USA<br />

and ChB: NA = 3.0; HE = 0.46). The lower genetic<br />

diversity found in the GMx, the largest manatee population<br />

in Mexico, is probably due to a combination of a founder<br />

effect, as this is the northern range of the sub-species of T.<br />

m. manatus, and a bottleneck event. The greater genetic<br />

diversity observed along the Caribbean coast, which also<br />

has the smallest estimated number of individuals, is possibly<br />

due to manatees that come from the GMx and Belize.<br />

There is evidence to support limited or unidirectional gene<br />

flow between these two important areas. The analyses<br />

presented here also suggest minimal evidence of a handful<br />

of individual migrants possibly between Florida and<br />

Mexico. To address management issues we suggest considering<br />

two distinct genetic populations in Mexico, one<br />

along the Caribbean coast and one in the riverine systems<br />

connected to the GMx.<br />

Keywords Antillean manatee Microsatellite<br />

Conservation genetic Genetic structure<br />

Introduction<br />

The National Manatee Conservation Program, implemented<br />

in 2001, divided Mexico’s manatee (Trichechus<br />

manatus manatus) population into three management units<br />

(MU) (SEMARNAT 2001) (Fig. 1). These MUs were<br />

based on the geographic distribution of the manatees, the<br />

different regional threats, local research capacities, regional<br />

socio-cultural factors, and similarities in habitat characteristics<br />

including water quality and availability that has<br />

led to a protection plan for the region (SEMARNAT 2001).<br />

An understanding of the manatee population structure and<br />

diversity in Mexico, based on population genetics, will<br />

provide a valuable tool to improve the scientific<br />

123


834 Genetica (2011) 139:833<strong>–</strong>842<br />

Fig. 1 Sampling sites, with<br />

existing manatee management<br />

units recognized in Mexico.<br />

Dots represent specific sampling<br />

sites<br />

information lending support for management actions, such<br />

as the new National Manatee Conservation Action Plan for<br />

Mexico (PACE manati). That plan is expected to be<br />

adopted this year (2011). In Mexico, manatees have been<br />

protected since 1922 (Diario Oficial 01/20/1922), listed as<br />

a threatened species in 1991 under Federal Law (SEDUE<br />

1991) and since 1994 has been classified as a species at risk<br />

of extinction by the Mexican Government (SEMARNAT<br />

2002). At the international level, Antillean manatees were<br />

classified as vulnerable by the <strong>International</strong> Union for<br />

Conservation of Nature (IUCN) Red List from 1982 to<br />

2007 and endangered since 2008 (Deutsch et al. 2008).<br />

West Indian manatees have also been listed as an endangered<br />

species by the Convention on <strong>International</strong> Trade in<br />

Endangered Species (CITES 2009) since 1975.<br />

The Antillean manatee occupies the tropical coastal<br />

waters of the Greater Antilles and Caribbean, extending from<br />

Mexico along Central and South America to Brazil (Husar<br />

1978; Lefebvre et al. 1989, 2001). The distribution of manatees<br />

in Mexico includes the coastal and wetland systems of<br />

the Gulf of Mexico (GMx) and the Caribbean coast (Fig. 1).<br />

The majority of manatees inhabit the wetland systems of<br />

Veracruz, Tabasco, Chiapas and Campeche states along the<br />

GMx, where a conservative estimate of population size is<br />

between 500 and 1,500 manatees (Olivera-Gómez 2006). It<br />

is estimated that approximately 200 to 250 manatees reside<br />

in Quintana Roo region (Morales-Vela and Padilla-Saldívar<br />

123<br />

2011) and much fewer reside within the Yucatan state<br />

(Morales-Vela et al. 2003). In Quintana Roo, most manatees<br />

occur in Chetumal Bay (ChB) with an estimate of about 100<br />

to 150 manatees (Morales-Vela and Padilla-Saldivar 2009a).<br />

In Ascencion Bay (AB) the estimate is smaller, with<br />

approximately 25 manatees counted during a single aerial<br />

survey in July of 2009 (Landero-Figueroa 2010). ChB and<br />

AB are located approximately 335 km apart (shore distance).<br />

North of the Yucatan Peninsula no estimates have<br />

been provided and manatee abundance is presumed to be<br />

very low (Morales-Vela et al. 2003).<br />

Previous genetic studies on the West Indian manatee<br />

suggest that their mitochondrial DNA (mtDNA) variability<br />

ranges between one (Florida) to as many as eight haplotypes<br />

(Colombia) per country (Garcia-Rodriguez et al.<br />

1998; Vianna et al. 2006). Genetic studies of manatees in<br />

Mexico, from Tabasco to Quintana Roo indicate differences<br />

among the populations. The biggest difference is<br />

between the GMx, where only one haplotype (J) is found,<br />

and the Caribbean coast where three haplotypes (J, A, A4)<br />

have been identified (Medrano-González et al. 1997;<br />

Castañeda-Sortibrán unpublished data). However, the small<br />

sample size from the GMx (n = 4) at the time of this<br />

mtDNA study was not sufficient to provide significant<br />

results.<br />

Historically, manatees were abundant in Mexico from<br />

Tamaulipas to the Yucatan Peninsula, but hunting for food


Genetica (2011) 139:833<strong>–</strong>842 835<br />

during the pre-Columbian period by Mayans (McKillop<br />

1985), the Spanish colonization and throughout the history<br />

of Mexico (Durand 1983), has resulted in the significantly<br />

reduced population occupying Mexico today. Even as<br />

recently as 1983, it was reported that manatees were occasionally<br />

harpooned in the region north of Quintana Roo as a<br />

source of food (Gallo-Reynoso 1983). Direct hunting is rare<br />

now in the GMx, but manatees caught in nets are still<br />

sometimes killed for food, with children reporting that they<br />

had eaten fresh manatee meat recently (Olivera-Gómez<br />

2006). Fishermen have also declared that the manatee<br />

population decline in the Northern and Western coasts of<br />

the Yucatan Peninsula was caused by a number of factors,<br />

including hunting for local consumption, entanglement as a<br />

result of higher net-fishing activities in rivers, and habitat<br />

destruction and coastal construction due to human population<br />

growth and hurricane impacts (Morales-Vela et al.<br />

2003). The last documented manatee hunting activities on<br />

the Northeast coast of the Yucatan Peninsula occurred in the<br />

1960<strong>–</strong>1970s, but since then opportunistic poaching has<br />

continued into the 1990s (Morales-Vela et al. 2003). In the<br />

GMx other factors such as natural or artificial closing of<br />

drainages from freshwater ecosystems have restricted<br />

manatee access to vegetation, increased water temperature,<br />

and have resulted in increases in pollution, sedimentation<br />

and eutrophication due to agricultural and poultry runoff<br />

(Olivera-Gómez 2006).<br />

The high level of anthropogenic and habitat destruction<br />

pressure on the manatee population may have caused a<br />

population bottleneck, which is known to reduce genetic<br />

diversity. Typically, habitat loss and degradation increase<br />

the rate of fragmentation resulting in the isolation of small<br />

populations and leads to an increased probability for<br />

inbreeding and unstable demographics (Frankham et al.<br />

2002).<br />

Herein, we present a fine-scale population structure of<br />

manatees in Mexico using analysis of microsatellite DNA<br />

markers. The genetic diversity and population structure<br />

were compared to a geographically close conspecific, the<br />

Florida manatee (Trichechus manatus latirostris), which<br />

has a larger estimated population size. The results of this<br />

microsatellite DNA marker study were compared with<br />

those obtained previously from an unpublished Mexican<br />

mtDNA study.<br />

Methods<br />

Microsatellite DNA amplification and fragment<br />

analysis<br />

We analyzed 94 samples from different regions of Mexico<br />

including the Peninsula of Yucatan MU (ChB: 51, AB: 15),<br />

the central GMx MU (Tabasco: 15, Chiapas: 5 and Campeche:<br />

1) and the north of the GMx MU (Veracruz: 7)<br />

(Fig. 1). Blood or skin tissue from the tail was collected<br />

from wild manatees captured for health assessment and<br />

radio tagging studies. Skin tissue was collected from carcasses<br />

recovered by manatee research projects throughout<br />

Mexico. Blood from captive manatees was also utilized for<br />

this study because their original rescue location was<br />

known. To resolve population structure, we included the<br />

genotypes of 95 individuals from Florida based on the<br />

estimated population size in each of the four recognized<br />

MU’s in that state (Pause 2007).<br />

Blood and tissue samples were preserved with lysis or<br />

tissue buffer respectively (lysis buffer: 100 mM Tris<strong>–</strong>HCl,<br />

100 mM EDTA, 10 mM NaCl, 1.0% SDS (White and<br />

Densmore 1992); SED tissue buffer: saturated NaCl;<br />

250 mM EDTA pH 7.5; 20% DMSO (Amos and Hoelzel<br />

1991; Proebstel et al. 1993)). DNA extractions, amplifications<br />

and fragment analysis were performed at the University<br />

of Florida, ICBR Genetic Analysis Laboratory in<br />

Gainesville, Florida, USA and at the US Geological Survey,<br />

Southeast Ecological Science Center Conservation<br />

Genetics Laboratory in Gainesville, FL, USA.<br />

DNA extractions were carried out using either a standard<br />

phenol<strong>–</strong>chloroform protocol (Hillis et al. 1990) or the<br />

DNeasy tissue extraction kit (QIAGEN, Valencia, CA, USA).<br />

Polymerase chain reaction (PCR) amplifications were<br />

performed for each sample for each of 13 microsatellite<br />

loci previously designed for manatees: TmaA02, TmaE02,<br />

TmaE08, TmaE11, TmaE26, TmaF14, TmaM79 (Garcia-<br />

Rodriguez et al. 2000), TmaSC13, TmaE7, TmaH13,<br />

TmaE14, TmaK01, TmaJ02 (Pause et al. 2007). Amplifications<br />

were performed in Biometra UNOII, UNO-<br />

Thermoblock, T-Gradient thermocyclers (BiometraÒ,<br />

Göttengen, Germany) or a PTC-200 (MJ Research, Waltham,<br />

MA) thermocycler using the following conditions:<br />

95°C for 5 min, 35 cycles of 95°C for 30 s, with the specific<br />

annealing temperature as listed in the original publication<br />

for each primer, with the exception of TmaM79 =<br />

54°C; TmaA02 = 56°C; TmaE02, TmaE11, TmaE26 and<br />

TmaF14 = 58°C; TmaE08 = 60°C for 30 s, 72°C for<br />

30 s, and a final extension at 72°C for 10 min. Amplifications<br />

were performed in a total volume of 15 lL, with<br />

10 ng target DNA, 19 Sigma PCR Buffer (10 mM Tris<strong>–</strong><br />

HCl, pH 8.3, 50 mM KCl, 0.001% gelatin), MgCl2 as<br />

indicated in the original publications, 0.2 mM each dNTP,<br />

0.04 units of Sigma JumpStart Taq polymerase (Sigma<strong>–</strong><br />

Aldrich, St. Louis, MO, USA), 0.25 lM each primer, and<br />

bovine serum albumin (BSA), where indicated (Garcia-<br />

Rodriguez et al. 2000; Pause et al. 2007). For fragment<br />

analysis, the forward primers were labeled with the fluorescent<br />

dyes HEX or 6-FAM for processing and visualization<br />

on an ABI 3730xl Automated DNA Analyzer.<br />

123


836 Genetica (2011) 139:833<strong>–</strong>842<br />

Fragment data from the PCR products were collected from<br />

the ABI 3730xl and analyzed using GENEMARKER 1.5<br />

(SoftGenetics 2008) to determine allele sizes. Allele sizes<br />

were standardized using previously analyzed Florida samples<br />

as the baseline.<br />

Data analysis<br />

CONVERT (Glaubitz 2004) was used to convert the data into<br />

different input file formats. We used STRUCTURE, version<br />

2.3.1, (Pritchard et al. 2000) to identify possible subpopulation<br />

designations, without an a priori assignment of the<br />

overall population structure in Mexico. The data from<br />

Mexico and Florida were analyzed together. The admixture<br />

model was used and the number of populations (K) was set<br />

from 1 to 10 with a burn-in period of 100,000 iterations,<br />

followed by 1,000,000 Monte Carlo Markov Chain iterations.<br />

Five independent analyses were simulated for each<br />

value of K. The value of K with the lowest posterior<br />

probability was identified as the optimum number of subpopulations,<br />

as recommended by the STRUCTURE manual.<br />

GENECLASS2 (Piry 2004) and WHICHRUN version 4.1 (Banks<br />

and Eichert 2000) were used to test individuals attributed to<br />

a different population than the geographic location<br />

assigned by STRUCTURE. GENALEX 6.2 (Peakall and Smouse<br />

2006) was used to calculate genetic distance between<br />

individuals which were used on a Principal Coordinate<br />

Analysis (PCA).<br />

Estimates of the effective population size were made by<br />

NEESTIMATOR using the Linkage Disequilibrium algorithm<br />

(Peel et al. 2004). GENALEX 6.2, GENEPOP 3.4 (Raymond<br />

and Rousset 1995; Rousset 2008) and ARLEQUIN 3.1 (Excoffier<br />

et al. 2005) were used to compare and determine the<br />

genetic diversity and genetic differentiation including<br />

allelic richness N A and N E, heterozygosity observed (HO)<br />

and expected (HE), estimate of population subdivision FST<br />

and RST and inbreeding coefficients (FIS). ARLEQUIN was<br />

used to check for deviation from Hardy<strong>–</strong>Weinberg equilibrium.<br />

GENEPOP webversion was used to examine for<br />

Linkage Disequilibrium. The presence of null alleles was<br />

analyzed with MICRO-CHECKER (Oosterhout et al. 2004). The<br />

presence of a potential bottleneck was estimated using<br />

BOTTLENECK examining the heterozygosity excess (Cornuet<br />

and Luikart 1996).<br />

Results<br />

Results from the three different software packages (GEN-<br />

ALEX, GENEPOP, and ARLEQUIN) were compared and similar<br />

results were observed for genetic diversity and population<br />

structure as had been previously published. Therefore, only<br />

the results from GENALEX are presented in Tables 1 and 3.<br />

123<br />

Population structure<br />

Results for the STRUCTURE analysis identified that k = 3<br />

was the appropriate number of clusters (Fig. 2) based on<br />

the manual’s recommendation of the lowest posterior<br />

probability LnP(D). Most individuals were assigned to a<br />

cluster with Q [ 80%. In all simulations, Florida manatees<br />

were assigned to a separate population cluster (92.8%<br />

assignment) and the GMx and ChB populations were<br />

assigned to a cluster that corresponded to their geography<br />

(89.0 and 84.6% respectively). Samples collected from AB<br />

were not as clearly delineated, and were clustered with<br />

either the GMx (56.1%) or ChB (41.6%).<br />

The STRUCTURE analysis suggests that there were some<br />

individuals that may have had mixed ancestry based on this<br />

analysis. Interestingly, some of the AB individuals with a<br />

high percentage of ancestry corresponding to the GMx<br />

cluster do not share the GMx’s unique haplotype. Instead,<br />

these individuals’ haplotypes corresponded to the AB/ChB<br />

regions. Four manatees sampled from ChB show high<br />

percent ancestry with the Florida cluster (between 59% and<br />

76%). One of them was a female manatee carcass sampled<br />

in ChB which shared a 59% ancestry with the Florida<br />

population. This manatee died of natural causes (birthing<br />

complications) and her calf was not recovered. Another<br />

was a female juvenile who shared 75.6% ancestry with<br />

Florida but her haplotype did not correspond with the<br />

Florida haplotype (Castañeda-Sortibrán pers. comm). One<br />

male manatee, who was rescued as a calf in March 2002<br />

and is now in captivity in Veracruz, also shares a high<br />

percent ancestry with Florida samples (51%). One female<br />

manatee from Veracruz shares ancestry with the GMx<br />

samples (44%) and ChB region samples (47%). Two<br />

manatees from Florida share a high percent ancestry (45<br />

and 60%) with the GMx individuals. Results from GENE-<br />

CLASS2 and WHICHRUN analysis corroborate these assignments<br />

to populations other than their geographic location<br />

(Table 1). The Veracruz samples have lower ancestry<br />

values and cannot be totally attributed to Florida or the<br />

GMx as they appear to be mixtures of both populations on<br />

GENECLASS2 and WHICHRUN analysis (Table 1).<br />

The pairwise FST value between ChB and AB is low but<br />

significant (FST = 0.047), and RST is not significant<br />

(RST = 0.016) (Table 2). All pairwise FST and RST values<br />

are presented in Table 2. Significant levels of subdivision<br />

were observed between the GMx and all other regions with<br />

FST and RST, and among all regions with FST. All values<br />

were significant when calculated between the three regions:<br />

the Caribbean coast (including ChB and AB), the GMx and<br />

Florida (data not show) but most of RST values were not<br />

significant when calculated by separating the Caribbean<br />

coast into two areas: ChB and AB and compared with the<br />

two other geographic regions: the GMx and Florida


Genetica (2011) 139:833<strong>–</strong>842 837<br />

Table 1 Population assignments of some individuals using STRUCTURE, GENECLASS2 and WHICHRUN<br />

Individual Sampling<br />

population<br />

STRUCTURE<br />

assignment<br />

(Table 2). A higher level of differentiation (via FST) was<br />

observed between ChB and the GMx than between ChB<br />

and Florida. This does not correspond to the STRUCTURE<br />

analysis, which consistently grouped Florida into its own<br />

cluster, separate from ChB. A handful of individuals had<br />

microsatellite genotypes that clustered with Florida, but<br />

overall, the group is distinct from Florida. This can also be<br />

STRUCTURE Q value GENECLASS2 results WHICHRUN<br />

BCh GMx FL Most Score Other Score Other Score<br />

Attribution<br />

probable for most population for population pop3<br />

population probable<br />

population<br />

(pop2) pop2 (pop3)<br />

BCH40 BCH BCh 0.98 0.01 0.02 BCH 100 FL 0 GMx 0 BCh<br />

BCH13 BCH BCh 0.97 0.02 0.01 BCH 100 GMx 0 FL 0 BCh<br />

BCH46 BCH FL 0.15 0.09 0.76 FL 81.1 BCH 15.44 GMx 3.46 FL<br />

BCH47 BCH FL 0.23 0.02 0.75 FL 85.32 BCH 14.35 GMx 0.33 FL<br />

BCH02 BCH FL 0.22 0.02 0.76 FL 85.16 BCH 14.69 GMx 0.14 FL<br />

N2 BCH FL 0.31 0.1 0.59 FL 76.96 BCH 15.55 GMx 7.49 FL<br />

TOO2 GMx GMx 0.01 0.98 0.01 GMx 99.96 BCH 0.05 FL 0 GMx<br />

V2 GMx GMx 0.01 0.98 0.01 GMx 99.73 BCH 0.27 FL 0 GMx<br />

V3 GMx BCh 0.47 0.43 0.1 GMx 90.13 BCH 9.86 FL 0.01 GMx<br />

V1 GMx FL 0.04 0.46 0.51 GMx 51.33 FL 47.97 BCH 0.7 GMx<br />

TM1007 FL FL 0.01 0.01 0.98 FL 100 GMx 0 BCH 0 FL<br />

TM772 FL FL 0.01 0.01 0.98 FL 100 BCH 0 GMx 0 FL<br />

TM775 FL FL 0.06 0.<strong>17</strong> 0.77 FL 79 GMx 18.55 BCH 2.45 FL<br />

TM925 FL FL 0.28 0.04 0.69 FL 76.06 BCH 23.35 GMx 0.59 FL<br />

TM523 FL FL 0.<strong>17</strong> 0.24 0.6 FL 78.25 GMx 15.36 BCH 6.39 GMx (similar<br />

probability<br />

for each<br />

population)<br />

TM636 FL FL 0.02 0.24 0.73 FL 86.96 GMx 12.14 BCH 0.9 FL<br />

TM993 FL GMx 0.05 0.64 0.31 GMx 57.76 FL 41.08 BCH 1.16 GMx<br />

TM641 FL FL 0.1 0.42 0.48 FL 63.68 GMx 27.06 BCH 9.26 FL<br />

Values generated by STRUCTURE in italics are individuals that have higher ancestry percentage associations when compared to other populations,<br />

relative to where the sample was collected. Underlined results indicate higher association to a population other than their original sampling<br />

location population. Results not in italics indicate two samples assigned to their geographic population for comparison<br />

Fig. 2 Proportions of ancestry for individuals were assessed without<br />

a priori information using Bayesian clustering via STRUCTURE. This<br />

graphic represents the best fit of the data, where three population<br />

clusters are clearly distinguished (K = 3). Manatees from Florida<br />

consistently grouped separately from the manatees from Mexico.<br />

Manatees from the Chetumal Bay (ChB) cluster together and those<br />

from the Gulf of Mexico (GMx) river systems clustered together.<br />

Samples from Ascencion Bay (AB) suggest it may be a mixing zone<br />

for the Mexican manatee populations<br />

observed in the PCA from genetic distance that shows<br />

separate clusters for Florida and Mexico (Fig. 3). The null<br />

hypothesis tested for heterozygosity excess using the<br />

Wilcoxon test, with BOTTLENECK software, provided<br />

(P = 0.0004 and P = 0.07) probabilities under Two-Phase<br />

mutation (TPM) and Step-wise mutation (SMM) respectively,<br />

with a normal L-shaped distribution for ChB<br />

123


838 Genetica (2011) 139:833<strong>–</strong>842<br />

Table 2 Pairwise F ST and R ST values comparing the Mexican manatee<br />

populations defined as Chetumal Bay (ChB), Ascencion Bay<br />

(AB) and the Gulf of Mexico (GMx), as well as the Florida manatee<br />

population, generated by GenAlEx 6.2<br />

ChB AB GMx Florida<br />

ChB <strong>–</strong> 0.016 0.114* 0.025<br />

AB 0.047* <strong>–</strong> 0.087* 0.014<br />

GMx 0.131* 0.088* <strong>–</strong> 0.089*<br />

Florida 0.096* 0.094* 0.106* <strong>–</strong><br />

* Significant values<br />

The FST values are presented below the diagonal, and RST values are<br />

above diagonal<br />

Fig. 3 PCA study from genetic distance using GENALEX. On the left<br />

is the Mexican population with individuals from the Caribbean and<br />

the Gulf of Mexico coasts. An admixture zone is apparent between<br />

these two areas. The population on the right corresponds to the<br />

Florida population<br />

population. For the GMx and AB populations the Wilcoxon<br />

test indicates a probability of 0.40 (TPM) and 0.<strong>17</strong> (SMM)<br />

with a shifted mode distribution for the GMx and 0.046<br />

(TPM) and 0.10 (SMM) with a shifted mode distribution<br />

for AB. For Florida, the Wilcoxon test shows a probability<br />

of 0.02 (TPM) and 0.55 (SMM) with a normal L-shaped<br />

distribution.<br />

Genetic diversity<br />

No evidence of null alleles was identified in either the<br />

Mexico or Florida populations. After 78 comparisons and a<br />

sequential Bonferroni correction, no linkage disequilibrium<br />

was observed (overall a = 0.05). All loci were in Hardy<strong>–</strong><br />

Weinberg equilibrium (HWE) after a sequential Bonferroni<br />

correction for all Mexican populations but TmaE02 was<br />

not within HWE for Florida. The error rate was determined<br />

by re-genotyping 12% of the samples. No errors were<br />

detected. Estimates of the mean number of alleles, effective<br />

number of alleles, and heterozygosity expected and<br />

observed, FIS, estimated population size and effective size<br />

for each determinate population (ChB, AB, the GMx and<br />

Florida) are presented in Table 3. Private alleles were<br />

found in Florida (n = 16), ChB (n = 2) and AB (n = 4).<br />

Discussion<br />

The GMx, with extended riverine and lagoon systems in<br />

Tabasco, Veracruz, Chiapas and Campeche, is a very<br />

favourable habitat for manatees. It has abundant vegetation,<br />

fresh water and areas where manatees can evade<br />

hunters. The Caribbean coast also represents very suitable<br />

habitat, especially ChB and AB, which are protected areas.<br />

ChB is the largest marine protected area at the state level in<br />

Mexico and is shared with Belize. AB is part of the Biosphere<br />

Reserve of Sian Ka’án, one of the largest Federal<br />

marine reserves in Mexico.<br />

Table 3 Diversity statistics over the 13 microsatellite loci examined for manatees from Mexico and Florida<br />

Population ChB AB GMx Florida<br />

Number of samples 51 15 28 95<br />

NA 3.00 ± 0.32 3.00 ± 0.32 2.62 ± 0.24 3.62 ± 0.48<br />

NE 1.99 ± 0.<strong>17</strong> 2.00 ± 0.22 1.84 ± 0.<strong>17</strong> 2.04 ± 0.16<br />

HE 0.46 ± 0.04 0.43 ± 0.05 0.41 ± 0.05 0.47 ± 0.04<br />

HO 0.47 ± 0.05 0.45 ± 0.07 0.44 ± 0.05 0.47 ± 0.04<br />

FIS -0.059 -0.035 -0.056 -0.006<br />

LD Abs Abs Abs Abs<br />

Null alleles Abs Abs Abs Abs<br />

Estimated population 100<strong>–</strong>150 23 (aerial survey) 500<strong>–</strong>1,500 5,000<br />

Effective population size 32.7 (23.9<strong>–</strong>47.4) 4.9 (4.0<strong>–</strong>6.2) 27.6 (18.0<strong>–</strong>49.4) 424.3 (208.3<strong>–</strong>6,816.5)<br />

The mean number of alleles (N A), effective number of alleles (N E), observed and expected heterozygosity (H O and H E, respectively), inbreeding<br />

coefficient (F IS), linkage disequilibrium after sequential Bonferroni correction (LD), null alleles, estimated population size, and effective<br />

population size for each population examined: Chetumal Bay (ChB), Ascencion Bay (AB), the Gulf of Mexico (GMx) and Florida for<br />

comparison<br />

123


Genetica (2011) 139:833<strong>–</strong>842 839<br />

Manatees are known to travel long distances along<br />

coastlines with suitable habitat (e.g. a Florida manatee was<br />

observed as far north as Rhode Island), as well as infrequent<br />

travel across open ocean (Alvarez-Alemán et al.<br />

2010; Deutsch et al. 2003; Fertl et al. 2005; Reep and<br />

Bonde 2006). Using existing tools, genetic analyses of<br />

Florida manatees have suggested little population structure<br />

throughout Florida (McClenaghan and O’Shea 1988; Garcia-Rodriguez<br />

et al. 1998; Pause pers. com.). Although<br />

manatees are capable of travelling along the shoreline<br />

between the GMx and the Caribbean coast, seasonal<br />

migration in response to cold weather is not necessary for<br />

survival in Mexico as it is in Florida.<br />

Radio tagging studies in ChB illustrated movement by<br />

five males and one female along the shoreline between<br />

discrete manatee habitats in ChB and Belize (Morales-Vela<br />

et al. 2007). One of these males was also observed participating<br />

in a mating group in Belize (Auil-Gomez pers.<br />

comm). The movement pattern from ChB to Belize was very<br />

similar as manatees made directed, continuous moves along<br />

the shoreline between discrete manatee habitats (Morales-<br />

Vela et al. 2007). None of the 19 manatees tagged with<br />

Argos-linked GPS radio tags, either from ChB or AB<br />

showed movements north of the initial tagging site (Morales-Vela<br />

and Padilla-Saldívar 2009b). The natural morphological<br />

features of ChB, oriented in the south of Mexico,<br />

easily connect ChB with both the Caribbean Sea and coastal<br />

Belize. An extensive coral reef barrier forms a natural<br />

marine corridor that may promote easier movement of the<br />

manatees between ChB to the south along the coast of<br />

Belize, than to the north of the Yucatan Peninsula. The<br />

Caribbean group is represented by ChB and AB, with a<br />

significant but low FST between these two locations. AB<br />

appears as a mixture of the southern portion of the Caribbean<br />

coast (ChB) and the GMx manatees, which is why the F ST<br />

between ChB and AB was significant but low (Fig. 2). It<br />

seems that migrants from the GMx breed in AB and do not<br />

continue south towards ChB. Mothers and calves are often<br />

observed in AB, which hosts suitable and abundant habitat<br />

for manatees. The mixture found in AB with the presence of<br />

individuals with both ancestral roots was not observed in the<br />

GMx region. The higher genetic diversity found in the<br />

Caribbean coast when compared to the GMx are consistent<br />

with results from the mtDNA data. Analyses of mtDNA<br />

analysis identified three haplotypes along the Caribbean<br />

coast, one common to Florida, another in common with<br />

Belize and the last shared a haplotype unique to the GMx<br />

(Castañeda-Sortibrán unpublished data). The Caribbean<br />

Cluster is geographically closer to Florida (approximately<br />

800 km) than to the GMx (approximately 1,150 km).<br />

A high level of genetic diversity in a population is<br />

thought to increase the probability of surviving a catastrophic<br />

event (Frankham et al. 2002). In this study, the<br />

mean number of observed alleles was very low and corresponds<br />

to less than what is predicted for wildlife populations<br />

that have been hunted or have been significantly fragmented<br />

(DiBattista 2008). However, these data describing low<br />

variability are consistent with other West Indian manatee<br />

populations throughout their range (Hunter et al. 2010,<br />

Pause pers. comm.). The lower genetic diversity, based on<br />

the number of alleles and heterozygosity found in the GMx,<br />

is probably due to a combination of a founder effect and a<br />

bottleneck event. This is likely due to manatees occupying<br />

the northern range of the subspecies and the extensive historical<br />

exploitation experienced during and since the pre-<br />

Columbian period up to the 1960<strong>–</strong>1970s. That depletion of<br />

manatee stocks resulted in the reduced population numbers<br />

observed today (Durand 1983; McKillop 1985; Medrano-<br />

González et al. 1997). The population with the greatest<br />

genetic diversity observed along the Caribbean coast also<br />

has the smallest estimated number of individuals (Table 3).<br />

This is possibly due to manatees that come from the GMx<br />

and Belize. Some breeding contribution of manatees from<br />

Belize is suggested by recent radio tagging data (Morales-<br />

Vela and Padilla-Saldívar 2009b) and by the presence of<br />

shared mtDNA haplotypes (Vianna et al. 2006). The proximity<br />

of the large Belize manatee population, only<br />

200<strong>–</strong>300 km south of ChB, may explain the dispersal route<br />

of the Caribbean coast manatees. Having a large population<br />

close by with an easy access route, makes it easier for the<br />

manatees to travel south than to try to make the longer<br />

journey to the GMx. There is evidence from radio tagging<br />

studies where some male manatees move from ChB to south<br />

Belize and return (Morales-Vela et al. 2007). However<br />

mtDNA shows different matrilineal trends in haplotype<br />

dispersal patterns in Belize and Mexico (Vianna et al. 2006)<br />

indicating that ChB and the Belize populations are not one<br />

well mixed population but have persisted independently<br />

over time. The genetic diversity along the Caribbean coast<br />

may have been recently influenced by potential migration<br />

from the growing manatee population in Florida as suggested<br />

by the few individuals that share remote ancestry<br />

with the Florida population.<br />

The Belize population shows a stronger separation from<br />

Florida (FST = 0.141) (Hunter et al. 2010) than the Mexican<br />

populations, with FST = 0.096 between the Mexican<br />

Caribbean coast and Florida and F ST = 0.106 between the<br />

GMx population and Florida. Belize and the Caribbean<br />

coast show a different pattern of haplotype distribution<br />

with two haplotypes in common and one private haplotype.<br />

The Florida haplotype, which is absent in Belize, yet is<br />

present in Mexico, confirms the closer relationship between<br />

Florida and Mexico than between Florida and Belize as<br />

confirmed by microsatellite data.<br />

Current collaborative studies utilizing samples from<br />

Mexico, Belize, Puerto Rico, Florida and other Caribbean<br />

123


840 Genetica (2011) 139:833<strong>–</strong>842<br />

countries will be completed in 2011<strong>–</strong><strong>2012</strong> to further assess<br />

the phylogeography and dispersal patterns of West Indian<br />

manatees. The effective population size is lower than the<br />

estimated population size for manatees in Mexico, which<br />

may be a reflection of the reduction in population size due<br />

to prior hunting pressure. The low Ne of the manatee<br />

population in the GMx may be due to a founder effect<br />

resulting in low genetic diversity, recent bottleneck events,<br />

and migratory limitations due to climate change and glacial<br />

periods. This could also be an artefact of the small sample<br />

size for the GMx. More samples are urgently needed from<br />

this area.<br />

Manatees in the GMx population tend to stay within the<br />

inland river systems, not in the coastal waters, and the<br />

group is likely fragmented by urban and agricultural<br />

development. An area of 140 km in the northeastern region<br />

of the Yucatan Peninsula is unsuitable habitat for manatees<br />

due to urban development, and the lack of vegetation and<br />

fresh water sources, resulting in a likely barrier to gene<br />

flow between the GMx and Caribbean coast (Medrano-<br />

González et al. 1997; Morales-Vela et al. 2003). Close to<br />

this zone, there are also significant coastal fishing activities<br />

and poaching (Morales-Vela et al. 2003). Examination of<br />

oceanographic structures shows a differentiation between<br />

the GMx and the Yucatan Peninsula (Laura Carrillo, pers.<br />

comm.; Merino 1997). This barrier may explain the low<br />

gene flow between the GMx and the Caribbean coast.<br />

Nevertheless, a group of six large manatees were observed<br />

in 2006 along the northeast coast of the Yucatan Peninsula<br />

(Reyes-Mendoza and Morales-Vela 2007). It is not known<br />

if these manatees were transient or resident.<br />

The analyses presented here suggest minimal evidence<br />

of a handful of individual migrants possibly between<br />

Florida and Mexico. One Florida manatee and six potential<br />

second generation migrants from Florida were detected in<br />

Mexico, which raises the question of whether migration or<br />

breeding occurs between the two subspecies located in<br />

Florida and Mexico. Two manatees from Florida share a<br />

high percent ancestry (45 and 60%) with the GMx individuals<br />

suggesting individuals may be second generation<br />

migrants. One of these manatees was a female with a<br />

known record of Florida maternal lineage from photoidentification.<br />

She was born in the Naples, FL area in 1995<br />

and captured in Port of the Islands, FL in February 2001.<br />

The other was the carcass of a male calf found in August<br />

1997 in the Banana River on the east coast of Florida<br />

(Beck, USGS, pers. comm.). One possibility is that<br />

migrants could be from Cuban waters; however, no genetic<br />

data are currently available from manatees in Cuba. Major<br />

currents go from Mexico to Florida and travel close to<br />

Cuba. Events such as tropical storms and hurricanes can<br />

change the currents and can separate manatees from the<br />

coast, taking them offshore and subject to drift (Langtimm<br />

123<br />

and Beck 2003). In that way, currents may also allow<br />

manatees to travel to Mexico from Florida. Manatees are<br />

able to survive short open water travels, and adults have no<br />

major natural predators so they will be able to survive<br />

incidental open sea travel. For example a known long-term<br />

resident female from the Florida manatee photo-identification<br />

catalog was sighted in Cuba in 2007 (Alvarez-Alemán<br />

et al. 2010). This manatee successfully travelled<br />

across open sea illustrating potential for population<br />

expansion. Reports of Florida manatees in the northern<br />

GMx indicate that they also can travel as far west as the<br />

southern Texas coast following the shoreline (Laguna<br />

Madre and Rio Grande) (Fertl et al. 2005). Historical distribution<br />

of manatees in Tamaulipas, Mexico which is<br />

north of the GMx shows a possible link that may exist<br />

between the GMx and Florida for manatees to travel<br />

(Lazcano-Barrero and Packard 1989). This leads to the<br />

question of successful reproduction between the two subspecies<br />

(T. m. manatus and T. m. latirostris) and the possible<br />

positive (increased variability, heterosis) or negative<br />

consequences (outbreeding depression) of breeding.<br />

Conclusion/conservation application<br />

The genetic structure of manatees in Mexico indicates two<br />

clusters with gene flow from the GMx to the Caribbean<br />

coast with no migration from the Caribbean back towards<br />

the GMx. This movement pattern could not be detected<br />

using mitochondrial DNA as there is only one haplotype<br />

(J) present in the GMx which is also present in the<br />

Caribbean. The three haplotypes found in the Caribbean are<br />

also found in the GMx, Florida and Belize. The individuals<br />

from AB appear as a mixture of the populations from ChB<br />

and the GMx. Additional evidence should be collected to<br />

determine the connectivity of other Central American<br />

populations of T. m. manatus, as well as the connectivity<br />

between the two subspecies.<br />

To address management issues, we suggest considering<br />

two distinct genetic populations in Mexico, one along the<br />

Caribbean coast and one in the riverine systems connected<br />

to the GMx. There is evidence to support unidirectional or<br />

limited gene flow between these two important areas.<br />

Special attention for conservation in the AB population<br />

would result in a healthier genetic stock, as this region is a<br />

source for potential mixture between the two primary<br />

genetic clusters.<br />

Based on this information it is important to maintain the<br />

natural migration routes of manatees from the GMx to the<br />

Caribbean area. This can be accomplished by conserving<br />

suitable manatee habitat along the north and east coasts of<br />

the Yucatan Peninsula, enforcing protection, reducing<br />

poaching and helping local initiatives that increase public


Genetica (2011) 139:833<strong>–</strong>842 841<br />

education. Ongoing tourism and urban development projects<br />

along those coasts are a major issue of concern.<br />

All of the captive manatees in Mexico were from the<br />

GMx region where they were originally rescued. However,<br />

most of them are in facilities located along the Caribbean<br />

coast. In the event of their release, returning them to the<br />

GMx would help maintain distinct populations. Also, if<br />

these facilities continue breeding manatees in captivity, a<br />

program that minimizes inbreeding and prevents inbred<br />

individuals or manatees with parents from different genetic<br />

populations from being released into the wild population<br />

must be considered. Genetic tools to assess lineages are<br />

currently being evaluated (Nourisson 2011).<br />

Acknowledgments This study is part of the doctoral dissertation<br />

research of the first author (CN), who was funded by a grant from the<br />

Secretaria de Relaciones Exteriores de México and by the Ministère<br />

des Affaires Etrangères Française—EGIDE. Financial support for<br />

training in Florida was provided by the Marine Mammal Commission.<br />

The project was funded by SEMARNAT/CONACYT (Project<br />

2002-C01-1128) and Dolphin Discovery for manatee captures and<br />

genetic analysis, as well as the US Geological Survey-Sirenia Project<br />

for genetic analysis. Mexican Government Research Permits and<br />

sample collection authorizations were obtained from DGVS # 03144;<br />

04513; 03670 and 03675. Samples were collected under scientific<br />

collection permits NUM/SGPA/DGVS/03144, NUM/SGPA/DGVS/<br />

04513, NUM/SGPA/DGVS/03670/06, SGPA/DGVS/04060/06;<br />

SGPA/DGVS/01103/07, SEDUM/SSMA/DGPPE/0191/2004 and<br />

RBSK OFICIO CUN 037/04. Samples were transported to the<br />

research laboratories in the US under CITES export permits<br />

MX22523, MX24463, MX28578, MX26485, MX38416, MX34645,<br />

MX31591 and CITES import permits 06US808447/9 and<br />

07US808447/9. Samples were processed in Florida under authority of<br />

USFWS wildlife research permit MA79<strong>17</strong>21 issued to the USGS<br />

Sirenia Project. All sample collection was done under approval of<br />

IACUC standards. Samples from the Gulf of Mexico area came from<br />

various projects and facilities: Dolphin Discovery, Veracruz Aquarium,<br />

Xcaret, ViaDelphi and Universidad Juárez Autónoma de<br />

Tabasco. We thank Dr. Margaret Hunter of the USGS Southeast<br />

Ecological Science Center for advice and help in the laboratory with<br />

various software packages. Any use of trade, product, or firm names is<br />

for descriptive purposes only and does not imply endorsement by the<br />

US Government.<br />

References<br />

Alvarez-Alemán A, Beck CA, Powell JA (2010) First report of a<br />

Florida manatee (Trichechus manatus latirostris) in Cuba. Aquat<br />

Mamm 36(2):148<strong>–</strong>153<br />

Amos B, Hoelzel AR (1991) Long-term preservation of whale skin<br />

for DNA analysis. Rep Int Whal Comm Special Issue 13:99<strong>–</strong>103<br />

Banks MA, Eichert W (2000) WHICHRUN (Version 3.2): a computer<br />

program for population assignment of individuals based on<br />

multilocus genotype data. J Hered 91:87<strong>–</strong>89<br />

CITES (2009) Appendices I, II and III. In: The CITES Appendices.<br />

Convention on <strong>International</strong> Trade in Endangered Species of<br />

Wild Fauna and Flora. http://www.cites.org/eng/app/appendices.<br />

shtml. Accessed 29 Oct 2009<br />

Cornuet JM, Luikart G (1996) Description and power analysis of two<br />

tests for inferring recent population bottlenecks from allele<br />

frequency data. Genetics 144:2001<strong>–</strong>2014<br />

Deutsch CJ, Reid JP, Bonde RK, Easton DE, Kochman HI, O’Shea TJ<br />

(2003) Seasonal movements, migratory behavior and site fidelity<br />

of West Indian manatees along the Atlantic Coast of the United<br />

States. Wildlife Monographs 151:1<strong>–</strong>77<br />

Deutsch CJ, Self-Sullivan C, Mignucci-Giannoni A (2008) Trichechus<br />

manatus. In: The IUCN Red List of Threatened Species<br />

Version 2009.1. <strong>International</strong> Union for Conservation of Nature<br />

and Natural Resources. http://www.iucnredlist.org/details/22<br />

103/0. Accessed 29 Oct 2009<br />

DiBattista JD (2008) Patterns of genetic variation in anthropogenically<br />

impacted populations. Conserv Genet 9(1):141<strong>–</strong>156<br />

Durand J (1983) Ocaso de sirenas, esplendor de manatíes, 2nd edn.<br />

Fondo de cultura económica, México<br />

Excoffier L, Laval G, Schneider S (2005) Arlequin ver. 3.0: an<br />

integrated software package for population genetics data analysis.<br />

Evol Bioinform Online 1:47<strong>–</strong>50<br />

Fertl DA, Schiro J, Regan GT, Beck CA, Adimey N, Price-May L,<br />

Amos A, Worthy GAJ, Crossland R (2005) Manatee occurrence<br />

in the northern Gulf of Mexico, West of Florida. Gulf Caribb Res<br />

<strong>17</strong>:69<strong>–</strong>94<br />

Frankham R, Ballou JD, Briscoe DA (2002) Introduction to conservation<br />

genetics. Cambridge University Press, Cambridge<br />

Gallo-Reynoso JP (1983) Notas sobre la distribución del manati<br />

(Trichechus manatus) en las costas de Quintana Roo. An Inst<br />

Biol Univ Nac Auton Mex 53(1982), Ser Zool (1):443<strong>–</strong>448<br />

Garcia-Rodriguez AI, Bowen BW, Domning D, Mignucci-Giannoni<br />

AA, Marmontel M, Montoya-Ospina RA, Morales-Vela B,<br />

Rudin M, Bonde RK, McGuire PM (1998) Phylogeography of<br />

the West Indian manatee (Trichechus manatus): how many<br />

populations and how many taxa? Mol Ecol 7:1137<strong>–</strong>1149<br />

Garcia-Rodriguez AI, Moraga-Amador D, Farmerie W, McGuire P,<br />

King TL (2000) Isolation and characterization of microsatellite<br />

DNA markers in the Florida manatee (Trichechus manatus<br />

latirostris) and their application in selected <strong>Sirenian</strong> species. Mol<br />

Ecol 9:2161<strong>–</strong>2163<br />

Glaubitz JC (2004) CONVERT: a user-friendly program to reformat<br />

diploid genotypic data for commonly used population genetic<br />

software packages. Mol Ecol Notes 4:309<strong>–</strong>310<br />

Hillis DM, Larson A, Davis SK, Zimmer EA (1990) Nucleic acids III:<br />

sequencing. In: Hillis DM, Moritz C (eds) Mol Syst. Sinauer<br />

Associates, Sunderland, pp 318<strong>–</strong>370<br />

Hunter ME, Auil-Gomez NE, Tucker KP, Bonde RK, Powell J,<br />

McGuire PM (2010) Low genetic variation and evidence of<br />

limited dispersal in the regionally important Belize manatee.<br />

Anim Conserv 13:592<strong>–</strong>602<br />

Husar LS (1978) Trichechus manatus. Mamm Sp 93:1<strong>–</strong>5<br />

Landero-Figueroa MM (2010) Distribución potencial del manatí<br />

(Trichechus manatus manatus) en Bahía de la Ascensión,<br />

Quintana Roo. Master dissertation. CINVESTAV-Merida,<br />

Mexico<br />

Langtimm CA, Beck CA (2003) Lower survival probabilities for adult<br />

Florida manatees in years with intense coastal storms. Ecol Appl<br />

13:257<strong>–</strong>268<br />

Lazcano-Barrero MA, Packard JM (1989) The ocurrence of manatees<br />

(Trichechus manatus) in Tamaulipas, Mexico. Mar Mamm Sci<br />

5(2):202<strong>–</strong>205<br />

Lefebvre LW, Marmontel M, Reid JP, Rathbun GB, Domning DP<br />

(2001) Status and biogeography of the West Indian manatee. In:<br />

Woods CA, Sergile FE (eds) Biogeography of the West Indies:<br />

patterns and perspectives, 2nd edn. CRC Press, Boca Raton,<br />

pp 425<strong>–</strong>474<br />

Lefebvre LW, O’Shea TJ, Rathbun GB, Best RC (1989) Distribution,<br />

status, and biogeography of the West Indian manatee. Biogeography<br />

of the West Indies:567<strong>–</strong>610<br />

McClenaghan LR, O’Shea TJ (1988) Genetic variability in the Florida<br />

manatee (Trichechus manatus). J Mamm 69:481<strong>–</strong>488<br />

123


842 Genetica (2011) 139:833<strong>–</strong>842<br />

McKillop HI (1985) Prehistoric exploitation of the manatee in the<br />

Maya and circum-Caribbean areas. World Archaeol 16:337<strong>–</strong>353<br />

Medrano-González L, Morales-Vela B, García-Rodríguez A, Robles-<br />

Saavedra R, Baker S (1997) Análisis preliminar de la variación<br />

del DNA mitocondrial y del complejo mayor de histocompatibilidad<br />

en la laguna de Catazajá, Chiapas, y en la bahía de<br />

Chetumal, Quintana Roo. In: Morales-Vela B, Medrano-González<br />

L (eds) Variación genética del manatí (Trichechus<br />

manatus), en el sureste de México y monitoreo con radiotransmisores<br />

en Quintana Roo. Informe final. El Colegio de la<br />

Frontera Sur, Comisión Nacional para el Conocimiento y Uso de<br />

la Biodiversidad (H164), Chetumal, pp 12<strong>–</strong>38. http://www.cona<br />

bio.gob.mx/institucion/proyectos/resultados/InfH164.pdf.pdf. Accessed<br />

29 Oct 2009<br />

Merino M (1997) Upwelling on the Yucatan Shelf: hydrographic<br />

evidence. J Mar Syst 13:101<strong>–</strong>121<br />

Morales-Vela B, Padilla-Saldivar JA (2009a) Aspectos biológicos de<br />

los manatíes en el sur de Quintana Roo. In: Espinoza-Ávalos J,<br />

Islebe GA, Hernández-Arana HA (eds) El sistema ecológico de la<br />

bahía de Chetumal/Corozal: costa occidental del Mar Caribe. El<br />

Colegio de la Frontera Sur. Chetumal, Quintana Roo, pp 115<strong>–</strong>123.<br />

http://w2.ecosur-qroo.mx/cna/julio/libbahia.pdf. Accessed 20 Feb<br />

2010<br />

Morales-Vela B, Padilla-Saldívar JA (2009b) Demografía, ecología y<br />

salud de la población de manatíes (Trichechus manatus manatus)<br />

en Quintana Roo, y su variación y representación genética en<br />

México. Informe Técnico Final Proyecto SEMARNAT/CONA-<br />

CYT 2002-C01-1128. El Colegio de la Frontera Sur, Chetumal,<br />

259 p<br />

Morales-Vela B, Padilla-Saldívar JA (2011) El manatí. La sirena del<br />

Caribe. In: Pozo C, Armijo-Canto N, Calmé S (eds) Riqueza<br />

biológica de Quintana Roo. Un análisis para su conservación.<br />

Tomo 1. ECOSUR, CONABIO, Gob. Edo. Q. R., PPD, México,<br />

pp 248<strong>–</strong>255<br />

Morales-Vela B, Padilla-Saldívar JA, Mignucci-Giannoni AA (2003)<br />

Status of the manatee (Trichechus manatus) along the northern<br />

and western coasts of the Yucatán Peninsula, México. Caribb J<br />

Sci 39(1):42<strong>–</strong>49<br />

Morales-Vela B, Padilla-Saldívar JA, Reid J, Butler S (2007) First<br />

records of long-distance manatee movements between Mexico<br />

and Belize. Sirenews 47:12<strong>–</strong>14<br />

Nourisson C (2011) Estructura genética de los manatíes en México.<br />

Tesis de doctorado, El Colegio de la Frontera Sur, Chetumal,<br />

Quintana Roo, México. 121pp<br />

Olivera-Gómez LD (2006) Estado actual del manatí (Trichechus<br />

manatus) en humedales del sur del Golfo de México. In: Primer<br />

Simposio para la Biología y Conservación del manatí antillano<br />

en Mesoamérica. Loma Linda University. http://resweb.llu.edu/<br />

rford/research/Manatees/ProgramayResumenes.pdf. Accessed 30<br />

Oct 2009<br />

Oosterhout CV, Hutchinson WF, Wills DPM, Shipley P (2004)<br />

MICRO-CHECKER: software for identifying and correcting<br />

genotyping errors in microsatellite data. Mol Ecol Notes 4:<br />

535<strong>–</strong>538<br />

Pause KC (2007) Conservation genetics of the Florida manatee,<br />

Trichechus manatus latirostris. PhD dissertation. University of<br />

Florida, USA<br />

Pause KC, Nourisson C, Clark A, Kellogg ME, Bonde RK, McGuire<br />

PM (2007) Polymorphic microsatellite DNA markers for the<br />

123<br />

Florida manatee (Trichechus manatus latirostris). Mol Ecol<br />

Notes 7:1073<strong>–</strong>1076<br />

Peakall R, Smouse PE (2006) Genalex 6: genetic analysis in Excel.<br />

Population genetic software for teaching and research. Mol Ecol<br />

Notes 6(1):288<strong>–</strong>295<br />

Peel D, Ovenden JR, Peel SL (2004) NeEstimator: software for<br />

estimating effective population size. In: Fisheries. Queensland<br />

Government, Department of Primary Industries and Fisheries.<br />

http://www.dpi.qld.gov.au/cps/rde/dpi/hs.xsl/28_6908_ENA_<br />

HTML.htm. Accessed 3 Nov 2009<br />

Piry S (2004) GeneClass2: A software for genetic assignment and<br />

first-generation migrant detection. J Her 95:536<strong>–</strong>539<br />

Pritchard JK, Stephens M, Donnelly P (2000) Inference of population<br />

structure using multilocus genotype data. Genetics 155:945<strong>–</strong>959<br />

Proebstel DS, Evans RP, Shiozawa DK, Williams RN (1993)<br />

Preservation of nonfrozen tissue samples from a salmonine fish<br />

Brachymystax lenok (Pallas) for DNA analysis. J Ichthyol 9:9<strong>–</strong><strong>17</strong><br />

Raymond M, Rousset F (1995) GENEPOP (version 1.2): population<br />

genetics software for exact tests and ecumenicism. J Hered<br />

86:248<strong>–</strong>249<br />

Reep RL, Bonde R (2006) The Florida manatee. The University Press<br />

of Florida, Gainesville, FL<br />

Reyes-Mendoza O, Morales-Vela B (2007) New observations of<br />

manatees off the northern coast of Quintana Roo, Mexico.<br />

Sirenews 47:14<br />

Rousset F (2008) Genepop’007: a complete reimplementation of the<br />

Genepop software for Windows and Linux. Mol Ecol Resour<br />

8:103<strong>–</strong>106<br />

SEDUE (1991) Acuerdo por el que se establecen los criterios<br />

ecológicos CT-CERN-001-91 que determinan las especies raras,<br />

amenazadas, en peligro de extinción o sujetas a protección<br />

especial y sus endemismos, de la flora y la fauna terrestres y<br />

acuáticas en la República Mexicana. In: Diario Oficial de la<br />

Federación, <strong>17</strong> de mayo de 1991. Gobierno Constitucional de los<br />

Estados Unidos Mexicanos. México, pp 7<strong>–</strong>36<br />

SEMARNAT (2001) Proyecto de conservación, recuperación y<br />

manejo del manatí Trichechus manatus en México. Ser PREP<br />

num 11. Secretaría de Medio Ambiente y Recursos Naturales,<br />

México<br />

SEMARNAT (2002) Norma Oficial Mexicana NOM-059-SEMAR-<br />

NAT-2001 Protección Ambiental—Especies nativas de México<br />

de Flora y Fauna Silvestres—Categorías de Riesgo y Especificaciones<br />

para su Inclusión, Exclusión o Cambio—Lista de<br />

Especies en Riesgo. In: Diario Oficial de la Federación, 6 de<br />

marzo de 2002, Segunda Sección. Gobierno Constitucional de<br />

los Estados Unidos Mexicanos. México, pp 1<strong>–</strong>80<br />

SoftGenetics (2008) GeneMarker, the biologist friendly software. In<br />

SoftGenetics, software powertools for genetic analysis. http://www.<br />

softgenetics.com/GeneMarker.html. Accessed 30 Oct 2009<br />

Vianna JA, Bonde RK, Caballero S, Giraldo JP, Lima RP, Clark A,<br />

Marmontel M, Morales-Vela B, de Souza MJ, Parr L, Rodríguez-<br />

López M, Mignucci-Giannoni AA, Powell JA, Santos FR (2006)<br />

Phylogeography, phylogeny and hybridization in trichechid<br />

sirenians: implications for manatee conservation. Mol Ecol<br />

15:433<strong>–</strong>447<br />

White PS, Densmore LD (1992) Mitochondrial DNA isolation. In:<br />

Hoezel AR (ed) Molecular genetic analysis of populations: a<br />

practical approach. IRL Press, Oxford University Press, New<br />

York, pp 29<strong>–</strong>58


Fish Sci (2011) 77:795<strong>–</strong>798<br />

DOI 10.1007/s12562-011-0388-x<br />

ORIGINAL ARTICLE Biology<br />

The differences in behavioral responses to a net obstacle<br />

between day and night in captive manatees; does entanglement<br />

happen at night?<br />

Mumi Kikuchi • Miwa Suzuki • Keiichi Ueda •<br />

Hirokazu Miyahara • Senzo Uchida<br />

Received: 23 March 2011 / Accepted: 27 June 2011 / Published online: 23 July 2011<br />

Ó The Japanese Society of Fisheries Science 2011<br />

Abstract Entanglement in fishing gear occurs in endangered<br />

manatees and may result in serious injury or death.<br />

Such incidents may happen more frequently at night when<br />

the animal’s visual sense is limited. In this study, we<br />

examined the differences in behavioral response of captive<br />

manatees to a net obstacle during light (day) and dark<br />

(night) periods. We used a plastic net as the obstacle, and<br />

video-recorded the manatees’ behavior. The experiments<br />

showed that captive manatees avoided the obstacle during<br />

the day more frequently than at night, which suggests that<br />

the manatees can perceive the obstacle more readily during<br />

light periods. However, there was no difference in the<br />

frequency of bumping or actively touching the obstacle<br />

between light and dark periods. The results suggest that the<br />

manatees can recognize the net obstacle even at night by<br />

purposely touching it, but they avoid it less frequently, and<br />

M. Kikuchi (&)<br />

Laboratory of Fisheries Biology, Graduate School<br />

of Agricultural and Life Science, University of Tokyo,<br />

1-1-1 Yayoi, Bunkyo, Tokyo 113-8657, Japan<br />

e-mail: mumi@cocoa.plala.or.jp<br />

M. Suzuki<br />

Department of Marine Science and Resources, College<br />

of Bioresource Sciences, Nihon University, 1866 Kameino,<br />

Fujisawa, Kanagawa 252-0880, Japan<br />

e-mail: miwa@brs.nihon-u.ac.jp<br />

K. Ueda H. Miyahara S. Uchida<br />

Okinawa Churaumi Aquarium, 424 Ishikawa, Motobu,<br />

Okinawa 905-0206, Japan<br />

e-mail: k_ueda@kaiyouhaku.or.jp<br />

H. Miyahara<br />

e-mail: h_miyahara@kaiyouhaku.or.jp<br />

S. Uchida<br />

e-mail: s_uchida@kaiyouhaku.or.jp<br />

that entanglement during light periods may occur during<br />

accidental bumping, rather than from a failure to recognize<br />

it altogether.<br />

Keywords Obstacle Behavior Trichechus manatus<br />

Introduction<br />

The greatest threat to many marine mammals is incidental<br />

entanglement and mortality in a fishing net, or bycatch [1].<br />

The three species of manatees are aquatic mammals of the<br />

order Sirenia, and are known as generalist aquatic herbivores<br />

with no apparent natural predators. However, there<br />

are conservation concerns for all three species because of<br />

collisions with boats, entanglement in fishing nets, hunting<br />

and habitat degradation. In Florida, collisions of the manatee<br />

Trichechus manatus latirostris with watercraft are the<br />

primary cause of human-contact-related manatee mortality,<br />

but other preventable deaths result from entrapment in<br />

water control structures and entanglement in fishing gear<br />

[2]. The Antillean manatee Trichechus manatus manatus is<br />

primarily threatened by entanglement in fishing nets [3, 4].<br />

The intensity of boat traffic throughout its range is relatively<br />

low in the areas used by Antillean manatees, and<br />

may not be a significant threat [3]. In the West African<br />

manatee Trichechus senegalensis, subsistence hunters have<br />

heavily exploited this species [5]; however, at present,<br />

incidental entanglement seems to be the most significant<br />

threat [6]. Even when manatees survive such encounters,<br />

entanglement may still result in serious injury, the loss of<br />

part of a pectoral flipper, and even death later on [5, 7, 8].<br />

A high degree of sensitivity for brightness discrimination<br />

has been observed in West Indian manatees, which is<br />

comparable to some piscivorous predators, such as fur seals<br />

123


796 Fish Sci (2011) 77:795<strong>–</strong>798<br />

[9]. Furthermore, manatees have dichromatic color vision<br />

that is most sensitive in the blue and green region of the<br />

spectrum [10, 11]. Color vision greatly enhances pattern<br />

recognition and may compensate for some deficiencies in<br />

resolution, at least in comparison to cetaceans with<br />

monochromatic vision [12]. Hartman [7] concluded that<br />

vision is the preferred means of exploring their surroundings<br />

in manatees. Therefore, we might safely assume that<br />

manatee entanglement happens more frequently at night<br />

when their visual sense may be more limited. In order to<br />

test this assumption, any possible difference in cognitive<br />

performance needs to be assessed between light and dark<br />

periods. In this study, we examined the difference in<br />

behavioral responses of captive manatees to a net obstacle<br />

between day and night periods, in order to better understand<br />

the apparent causes of manatee entanglement.<br />

Materials and methods<br />

Experiments were conducted with four captive Antillean<br />

manatees Trichechus manatus manatus from June to September<br />

2004 at the Okinawa Churaumi Aquarium, Japan<br />

(Table 1). The manatees were kept in two indoor pools<br />

with the sexes separated. The first pool, containing two<br />

males, was 8.0 m long, 4.9 m wide and 3.0 m deep; the<br />

second pool, containing two females, was 8.0 m long,<br />

10.8 m wide and 3.0 m deep. The indoor facility has large<br />

windows that allow natural sunlight into the pools.<br />

Experiments were carried out randomly in the morning<br />

(0730 hours) and at night (1900<strong>–</strong>2000 hours). All experiments<br />

were conducted outside of regular business hours to<br />

minimize potential human interference. A total of eight and<br />

nine independent trials were conducted during the daytime<br />

and nighttime, respectively. In each trial, the subject’s<br />

behaviors were recorded for one hour with a video camera<br />

(FV40/FV300, Canon, Japan). During nighttime trials, a<br />

night-vision scope (WV-BP50, Panasonic, Japan) was also<br />

used to observe the subject’s behavior. Illumination<br />

intensity was recorded near the surface of the water using<br />

a digital luxmeter (LX-1332, Custom Co., Japan) at the<br />

beginning and end of each trial.<br />

Table 1 Name, standard body length (SL), body mass (BM), age<br />

(estimated for all except Yuma), and sex of each manatee used in the<br />

current study (data from 2004)<br />

ID Name SL (m) BM (kg) Age (year) Sex<br />

1 Yukatan 2.7 302 26 Male<br />

2 Ryu 2.7 3<strong>17</strong> 14 Male<br />

3 Maya 2.8 412 <strong>17</strong> Female<br />

4 Yuma No data No data 3 a<br />

Female<br />

a Yuma was born at the aquarium<br />

123<br />

Fig. 1 Photo of the black plastic net obstacle used in this study. The<br />

obstacle was constructed from PVC pipe and plastic netting, with<br />

weights attached to the bottom<br />

In order to examine the manatee’s behavioral response<br />

to the obstacle, a plastic net with no danger of entanglement<br />

was used. The obstacle (150 9 160 cm, black) was<br />

constructed from PVC tubing and plastic netting (1.5 9<br />

1.5 cm mesh size, 0.2 cm width, Takiron Co., Ltd.), with<br />

weights attached to the lower part of the obstacle to submerge<br />

it perpendicularly in the water (Fig. 1). The obstacle<br />

was placed in the pool such that the manatees had sufficient<br />

space to pass around it with no interference. The position<br />

of the obstacle was changed for each trial. To analyze the<br />

behavioral response to the obstacle, each individual’s<br />

behavior was classified into three categories: avoiding,<br />

bumping, or active touching with their snout or pectoral fin.<br />

The number of times these behaviors were observed and<br />

the duration of active touching were recorded. In addition,<br />

the number and durations of events of inactive behavior at<br />

the surface or on the bottom were calculated in order to<br />

compare with their active behaviors between day and night.<br />

To confirm the difference in illumination intensity<br />

between day and night, a generalized linear mixed model<br />

(GLMM) was analyzed with gamma errors, with each trial<br />

considered a random factor, and day<strong>–</strong>night differences<br />

treated as explanatory variables. To test the effect of the<br />

day<strong>–</strong>night difference on the number and duration of each<br />

recorded behavior, a GLMM was used with Poisson and<br />

gamma errors, respectively, and each individual manatee


Fish Sci (2011) 77:795<strong>–</strong>798 797<br />

was included as a random factor. The most parsimonious<br />

model was selected based on the Akaike information criterion<br />

(AIC). For all statistical analyses, we used the<br />

software R, and the GLMM was analyzed using the<br />

package lme4 and the function lmer.<br />

Results<br />

The GLMM for illumination intensity including the day<strong>–</strong><br />

night difference as an explanatory variable was the best<br />

model, with the lowest AIC. The model had an AIC value<br />

that was 578.2 lower than the second best model, which<br />

included random effects only. This suggests that illumination<br />

intensity clearly differed between day and night<br />

(GLMM: day<strong>–</strong>night, estimated value during the day:<br />

1161.0 ± 63.5 SE, night: 1.1 ± 63.5 SE).<br />

The GLMM for the number of events of avoiding<br />

behavior that included the day<strong>–</strong>night difference as an<br />

explanatory variable was the best model, with the lowest<br />

AIC (Table 2). This model had an AIC value that was 31.1<br />

lower than the second best model, which included random<br />

effects only. These results suggest that the number of<br />

events of avoiding behavior differed between the day and<br />

night: these subjects showed a greater number of events of<br />

obstacle avoidance behavior during the day than at night<br />

(GLMM: day<strong>–</strong>night, estimated value during the day:<br />

1.7 ± 3.5 SE, at night: 0.8 ± 1.1 SE).<br />

The model with the lowest AIC value with respect to<br />

the number of events of bumping or active touching<br />

included random effects only (Table 2). The model had an<br />

AIC value of 0.8 and 1.7 lower than the second best<br />

model, which included the day<strong>–</strong>night difference. The<br />

GLMM for the number of events of bumping (GLMM:<br />

random effects only, 0.9 ± 1.3 SE) and active touching<br />

(GLMM: random effects only, 7.2 ± 2.1 SE) revealed<br />

that it was not related to the day<strong>–</strong>night difference. The<br />

GLMM for the duration of active touching, including<br />

random effects only, was the best model with the lowest<br />

AIC. The AIC value was 2.0 lower than the second best<br />

model, which included the day<strong>–</strong>night difference. These<br />

results suggest that the duration of active touching was not<br />

different between day and night (GLMM: random effects<br />

only, 49.8 ± 12.8 SE).<br />

The GLMM for the number or duration of events of<br />

inactive behavior, including random effects only, was the<br />

best model with the lowest AIC. The model had an AIC<br />

value of 0.7 and 0.8 lower than the second best model,<br />

which included day<strong>–</strong>night. These results indicate that the<br />

number (GLMM: random effects only, 1.3 ± 1.5 SE) and<br />

duration (GLMM: random effects only, 104.4 ± 10.7 SE)<br />

of events of inactive behavior were not related to the day<strong>–</strong><br />

night difference.<br />

Discussion<br />

In this study, the number and duration of events of<br />

inactive behavior did not differ between day and night.<br />

Therefore, we concluded that our subjects’ activities did<br />

not differ between day and night, and that it did not affect<br />

their behavioral responses to the obstacle during the<br />

experiments.<br />

Focusing on the differences in behavioral responses to a<br />

net obstacle between day and night, we found that captive<br />

manatees show a greater number of avoidance behaviors<br />

during the day than at night. These results suggest that<br />

manatees can recognize the obstacle more easily during<br />

light periods. However, there was no difference in the<br />

number of bumping events between light and dark periods.<br />

Therefore, accidental bumping may be expected to occur<br />

regardless of the level of illumination and manatee’s higher<br />

degree of associated recognition.<br />

All sirenian species have sparse hairs on their bodies and<br />

orofacial that are of the sinus type, and typically tactile in<br />

function [13<strong>–</strong>18]. In fact, it has been specifically reported<br />

that a manatee’s orofacial hairs are used for tactile exploration<br />

and discrimination [13<strong>–</strong>16, 18]. In the current study,<br />

the number and the duration of active touching behavioral<br />

events did not differ between day and night. These results<br />

indicate that manatees heavily rely on their tactile senses to<br />

recognize an obstacle, even during light periods that<br />

facilitate purely visual recognition. In field studies, when<br />

wild Antillean manatees are caught by encircling them with<br />

Table 2 The GLMM models for the number of events of each behavioral response to the obstacle (avoidance, active touching, and bumping of<br />

the obstacle) along with AIC and delta AIC (DAIC) values<br />

Model The number of events of each behavioral response to the obstacle<br />

Avoidance Active touching Bumping<br />

AIC DAIC AIC DAIC AIC DAIC<br />

Random effects only 200.7 31.1 180.8 0.0 48.8 0.0<br />

Day<strong>–</strong>night 169.6 0.0 182.5 1.7 49.6 0.8<br />

Favoured models (bold type) were evaluated on the basis of the lowest AIC<br />

123


798 Fish Sci (2011) 77:795<strong>–</strong>798<br />

nets, the animals often explore the ‘‘weak areas’’ of the net<br />

via tactile contact in order to escape (Olivera-Gomez LD,<br />

2011, pers. comm.). They tend to descend to the bottom of<br />

the net and pull it up with its snout. There are many reports<br />

of these escape behaviors by local fishermen. Wild Antillean<br />

manatees can escape from the net in this way, and this<br />

more often occurs at deeper depths, where the net is more<br />

vertical (Olivera-Gomez LD, 2011, pers. comm.). In shallow<br />

depths, where the net forms a bag, manatees enter the<br />

fold of the net and are more easily caught on it (Olivera-<br />

Gomez LD, 2011, pers. comm.). In addition, entanglement<br />

often occurs with larger mesh size nets, which are used to<br />

catch large fishes (Olivera-Gomez LD, 2011, pers. comm.).<br />

Therefore, we further recommend examining the position<br />

of the net relative to the depth and the effect of the mesh<br />

size in order to clarify the causes of manatee entanglement.<br />

Acknowledgments We would like to thank all of the staff at the<br />

Okinawa Churaumi Aquarium for their assistance in conducting these<br />

experiments. We would like to thank K. Asahina at Nihon University,<br />

H. Kato at Tokyo University of Marine Science and Technology, L.<br />

D. Olivera-Gomez at the Autonomous University Juarez of Tabasco,<br />

and D. Gonzalez-Socoloske at Duke University for their discussions<br />

and insightful comments on previous drafts. We also thank J.<br />

A. Mobley and D. Gonzalez-Socoloske for correcting the English<br />

manuscript. The present study was supported by the Fisheries<br />

Research Agency, National Research Institute of Far Seas Fisheries.<br />

References<br />

1. Grech A, Marsh H, Coles R (2008) A spatial assessment of the<br />

risk to a mobile marine mammal from bycatch. Aquat Conserv<br />

Mar Freshw Ecosys 18:1127<strong>–</strong>1139<br />

2. Reep RL, Bonde RK (2006) The Florida manatee: biology and<br />

conservation. University Press of Florida, Gainesville<br />

3. de Thoisy B, Spiegelberger T, Rousseau S, Talvy G, Vogel I, Vie<br />

JC (2003) Distribution, habitat, and conservation status of the<br />

West Indian manatee Trichechus manatus in French Guiana.<br />

Oryx 37:431<strong>–</strong>436<br />

4. Castelblanco-Martinez DN, Bermudez-Romero AL, Gomez-<br />

Camelo IV, Rosas FCW, Trujillo F, Zerda-Ordonez E (2009)<br />

Seasonality of habitat use, mortality and reproduction of the<br />

123<br />

vulnerable Antillean manatee Trichechus manatus manatus in the<br />

Orinoco River, Colombia: implications for conservation. Oryx<br />

43:235<strong>–</strong>242<br />

5. Reeves RR, Tuboku-Metzger D, Kapindi RA (1988) Distribution<br />

and exploitation of manatees in Sierra Leone West Africa. Oryx<br />

22:75<strong>–</strong>84<br />

6. Silva MA, Araujo A (2001) Distribution and current status of<br />

the West African manatee (Trichechus senegalensis) in Guinea-<br />

Bissau. Mar Mammal Sci <strong>17</strong>:418<strong>–</strong>424<br />

7. Hartman DS (1979) Ecology and behavior of the manatee<br />

(Trichechus manatus) (Spec Publ 5). Am Soc Mammals,<br />

Lawrence<br />

8. Beck CA, Barros NB (1991) The impact of debris on the Florida<br />

manatee. Mar Pollut Bull 22:508<strong>–</strong>510<br />

9. Griebel U, Schmid A (1997) Brightness discrimination ability in<br />

the West Indian manatee (Trichechus manatus). J Exp Biol<br />

200:1587<strong>–</strong>1592<br />

10. Cohen JL, Tucker GS, Odell DK (1982) The photoreceptors of<br />

the West Indian manatee. J Morphol <strong>17</strong>3:197<strong>–</strong>202<br />

11. Griebel U, Schmid A (1996) Color vision in the manatee<br />

(Trichechus manatus). Vision Res 36:2747<strong>–</strong>2757<br />

12. Bauer GB, Colbert DE, Gaspard JCIII, Littlefield B (2003)<br />

Underwater visual acuity of Florida manatees (Trichechus manatus<br />

latirostris). Int J Comp Sociol Psychol 16:130<strong>–</strong>142<br />

13. Marshall CD, Clark LA, Reep RL (1998) The muscular hydrostat<br />

of the Florida manatee (Trichechus manatus latirostris): a functional<br />

morphological model of perioral bristle use. Mar Mamm<br />

Sci 14:290<strong>–</strong>303<br />

14. Reep RL, Marshall CD, Stoll ML, Whitaker DM (1998) Distribution<br />

and innervation of facial bristles and hairs in the Florida<br />

manatee (Trichechus manatus latirostris). Mar Mamm Sci<br />

14:257<strong>–</strong>273<br />

15. Bachteler D, Dehnhardt G (1999) Active touch performance in<br />

the Antillean manatee: evidence for a functional differentiation of<br />

facial tactile hairs. Zool 102:61<strong>–</strong>69<br />

16. Reep RL, Stoll ML, Marshall CD, Homer BL, Samuelson DA<br />

(2001) Microanatomy of facial vibrissae in the Florida manatee:<br />

the basis for specialized sensory function and oripulation. Brain<br />

Behav Evol 58:1<strong>–</strong>14<br />

<strong>17</strong>. Reep RL, Marshall CD, Stoll ML (2002) Tactile hairs on the<br />

postcranial body in Florida manatees: a mammalian lateral line?<br />

Brain Behav Evol 59:141<strong>–</strong>154<br />

18. Marshall CD, Maeda H, Iwata M, Furuta M, Asano S, Rosas F,<br />

Reep RL (2003) Orofacial morphology and feeding behaviour of<br />

the dugong, Amazonian, West African and Antillean manatees<br />

(Mammalia: Sirenia): functional morphology of the muscular<strong>–</strong><br />

vibrissal complex. J Zool 259:245<strong>–</strong>260


Low genetic variation and evidence of limited dispersal in<br />

the regionally important Belize manatee<br />

M. E. Hunter 1,2 , N. E. Auil-Gomez 3,4 , K. P. Tucker 5 , R. K. Bonde 1,2 , J. Powell 4 & P. M. McGuire 2<br />

1 Sirenia Project, Southeast Ecological Science Center, U.S. Geological Survey, Gainesville, FL, USA<br />

2 Department of Physiological Sciences, College of Veterinary Medicine, University of Florida,Gainesville, FL, USA<br />

3 Wildlife Trust, Belize City, Belize<br />

4 Sea to Shore Alliance, St Petersburg, FL, USA<br />

5 College of Marine Science, University of South Florida, St. Petersburg, FL, USA<br />

Keywords<br />

conservation genetics; low diversity;<br />

microsatellite; mitochondria; marine<br />

mammal; West Indian manatee.<br />

Correspondence<br />

Margaret E. Hunter, Sirenia Project,<br />

Southeast Ecological Science Center, US<br />

Geological Survey, 2201 NW 40th Terrace,<br />

Gainesville, FL 32605, USA. Tel: +1 352<br />

264 3484; Fax: +1 352 378 4956<br />

Email: mkellogg@usgs.gov<br />

Received 26 October 2009; accepted 13 June<br />

2010<br />

doi:10.1111/j.1469-<strong>17</strong>95.2010.00383.x<br />

Introduction<br />

Abstract<br />

The West Indian manatee Trichechus manatus is a threatened<br />

aquatic mammal found throughout the south-eastern United<br />

States, Central and South America and the Caribbean. The<br />

Florida manatee Trichechus manatus latirostris and Antillean<br />

manatee Trichechus manatus manatus are the two recognized<br />

subspecies of the West Indian manatee. A low reproductive<br />

rate, environmental impacts and direct threats from the human<br />

population have historically limited manatee population<br />

growth. In 1982, the <strong>International</strong> Union for Conservation of<br />

Nature classified all West Indian manatee populations as<br />

vulnerable to extinction (IUCN, 2007).<br />

During the <strong>17</strong>th through 19th centuries, the Spaniards in<br />

Belize and throughout the region severely exploited the<br />

592<br />

Animal Conservation. Print ISSN 1367-9430<br />

The Antillean subspecies of the West Indian manatee Trichechus manatus is found<br />

throughout Central and South America and the Caribbean. Because of severe<br />

hunting pressure during the <strong>17</strong>th through 19th centuries, only small populations of<br />

the once widespread aquatic mammal remain. Fortunately, protections in Belize<br />

reduced hunting in the 1930s and allowed the country’s manatee population to<br />

become the largest breeding population in the Wider Caribbean. However,<br />

increasing and emerging anthropogenic threats such as coastal development,<br />

pollution, watercraft collision and net entanglement represent challenges to this<br />

ecologically important population. To inform conservation and management<br />

decisions, a comprehensive molecular investigation of the genetic diversity,<br />

relatedness and population structure of the Belize manatee population was<br />

conducted using mitochondrial and microsatellite DNA. Compared with other<br />

mammal populations, a low degree of genetic diversity was detected (HE=0.455;<br />

N A=3.4), corresponding to the small population size and long-term exploitation.<br />

Manatees from the Belize City Cayes and Southern Lagoon system were genetically<br />

different, with microsatellite and mitochondrial FST values of 0.029 and<br />

0.078, respectively (P 0.05). This, along with the distinct habitats and threats,<br />

indicates that separate protection of these two groups would best preserve the<br />

region’s diversity. The Belize population and Florida subspecies appear to be<br />

unrelated with microsatellite and mitochondrial F ST values of 0.141 and 0.63,<br />

respectively (P 0.001), supporting the subspecies designations and suggesting<br />

low vagility throughout the northern Caribbean habitat. Further monitoring and<br />

protection may allow an increase in the Belize manatee genetic diversity and<br />

population size. A large and expanding Belize population could potentially assist<br />

in the recovery of other threatened or functionally extinct Central American<br />

Antillean manatee populations.<br />

Antillean subspecies for sustenance (Lefebvre et al., 2001).<br />

By 1936, the population decline was so severe in Belize that<br />

Manatee Protection Ordinances were introduced to preserve<br />

the population (McCarthy, 1986). Today, the Belize manatee<br />

is listed as endangered by the Belize Wildlife Protection<br />

Act of 1981, Part II, Section 3(a) (Auil, 1998, 2004). A<br />

Manatee Recovery Plan was also proposed requesting<br />

information on habitat use and movement patterns to aid<br />

the development of conservation policies for the protection<br />

of the Belize manatee (Auil, 1998).<br />

While studying Caribbean manatee populations, O’Shea<br />

& Salisbury (1991) concluded that ‘Belize remains one of the<br />

last strongholds for the species in this part of the world.’<br />

From 1977 to 1991, the Belize manatee population size<br />

appeared stable, however, an overall negative trend in<br />

Animal Conservation 13 (2010) 592<strong>–</strong>602 c 2010 The Authors. Animal Conservation c 2010 The Zoological Society of London


M. E. Hunter et al.<br />

numbers was detected from 1997 to 2002 (Auil, 2004).<br />

Although today Belize is considered the largest breeding<br />

Antillean population, it is still under pressure from anthropogenic<br />

threats and the census size of 1000 individuals is<br />

well below that recommended for the long-term genetic<br />

sustainability of a population (Frankham, Ballou & Briscoe,<br />

2002; Quintana-Rizzo & Reynolds III, 2007).<br />

Little is known of the abundance or distribution of<br />

manatees in the other Central American populations,<br />

although they are considered elusive and thought to be rare<br />

(Quintana-Rizzo & Reynolds III, 2007). However, supplementation<br />

from Belize or Mexico manatees in concert with<br />

coordinated regional efforts could potentially assist in the<br />

recovery of the Guatemala populations and those further<br />

the south. The large Mexico population is also protected<br />

Reduced Belize manatee dispersal and genetic variation<br />

Figure 1 Geographic map of Belize with the<br />

three study sites inset, Belize City Cayes<br />

(north), Northern & Southern Lagoons (center),<br />

and Placencia Lagoon (south).<br />

and manatees from Mexico have been documented traveling<br />

to Belize and even participating in a mating herd (Auil et al.,<br />

2007; Morales-Vela et al., 2007). Quintana-Rizzo & Reynolds<br />

III (2007) suggest that safeguarding the Belize and<br />

Mexico populations as a resource of regional importance<br />

should be a high conservation priority.<br />

Long-term exploitation and small population sizes<br />

can lower genetic diversity and decrease fitness-related<br />

traits, such as fecundity and survival (Hoelzel et al., 1993;<br />

Frankham et al., 2002; Dixon et al., 2007). Genetic variation<br />

often acts in concert with demographic stochasticity in<br />

small populations. Loss of diversity can ultimately lead to<br />

an ‘extinction vortex,’ or a cyclic reduction in population<br />

size, resulting in loss of the population (Gilpin & Soulé,<br />

1986).<br />

Animal Conservation 13 (2010) 592<strong>–</strong>602 c 2010 The Authors. Animal Conservation c 2010 The Zoological Society of London 593


Reduced Belize manatee dispersal and genetic variation M. E. Hunter et al.<br />

A previous mitochondrial DNA (mtDNA) study identified<br />

three haplotypes (A03, A04 and J01) in 43 animals in Belize<br />

(Vianna et al., 2006). The resultant haplotype parameters<br />

included 28 polymorphic sites and nucleotide substitutions<br />

with h=0.558 and p=0.037. The distribution and genetic<br />

structure of the mtDNA haplotypes was not addressed, but<br />

could provide valuable information on the population structure,<br />

especially coupled with biparentally inherited nuclear<br />

data. Presented here are mitochondrial and nuclear microsatellite<br />

DNA studies addressing the genetic diversity,<br />

relatedness and population structure in Belize manatees.<br />

The Antillean Belize population is also compared with<br />

T. m. latirostris to provide information on the extent of West<br />

Indian manatee migration and relatedness.<br />

Materials and methods<br />

Sample collection and DNA extraction<br />

Manatee epidermis tissue and/or white blood cells were<br />

collected from recovered carcasses or during wild manatee<br />

health assessments in Belize from 1997 to 2007. Genetic<br />

samples were collected from a total of 118 unique animals<br />

in 213 health assessments (50% female and 50% male; Auil<br />

et al., 2007) and 12 carcass samples. Genomic DNA was<br />

isolated using Qiagen’s DNeasy blood and tissue kits<br />

(Valencia, CA, USA). Within the Florida dataset, 96 genotypes<br />

were randomly chosen, proportionally representing<br />

the four demographically defined management units<br />

(K.P. Tucker et al., unpubl. data).<br />

Study locations<br />

The Belize City Cayes (BCC), including the near shore and<br />

outlying cayes, are utilized by the tourist industry and experience<br />

heavy boat traffic and recreational activities from Belize<br />

City, the largest port in the country (Fig. 1). Manatee tours are<br />

largely conducted around Swallow Caye within the BCC.<br />

The Southern Lagoon system (SLS) consists of the Southern<br />

and Northern Lagoons and has the highest recorded<br />

number of manatees, with 55 individuals sighted in a single<br />

aerial survey (O’Shea & Salisbury, 1991). While the population<br />

is not heavily affected by watercraft, it is potentially<br />

impacted by salinity changes due to increased rainfall,<br />

influencing the abundance of aquatic vegetation, the primary<br />

food source for manatees.<br />

Placencia Lagoon (PL) is located 75 miles south of the<br />

SLS (Fig. 1). Placencia Village is a popular tourist destination,<br />

with a large amount of coastal development and<br />

agricultural runoff severely altering the environment and<br />

aquatic vegetation. Nitrification from shrimp farm effluent<br />

causes extensive algae blooms that can compromise the<br />

quality and growth of seagrass (Thornton, Shanahan &<br />

Shanahan, 2003). Nitrification also provides a food source<br />

for a marine snail, thought to be an intermediate host of a<br />

respiratory fluke Pulmonicola cochleotrema (Blair, 2005),<br />

which is known to parasitize manatee lungs (Beck &<br />

Forrester, 1988). Manatees in PL present at necropsy with<br />

594<br />

large loads of P. cochleotrema and copious mucoid discharge<br />

that could hinder respiration (Auil et al., 2007).<br />

The carcass samples utilized here were recovered from<br />

Corazol (n=1), Stan Creek (n=1), the Belize River mouth<br />

in Belize City (an area of high watercraft activity), or south of<br />

the river, possibly caught in the ocean current (n=10). The<br />

northern most manatee reserve in Belize is a section of<br />

Chetumal Bay, which continues into Mexico (Fig. 1). Samples<br />

collected in Chetumal Bay are not included in this study.<br />

Laboratory procedures<br />

Mitochondrial DNA analysis<br />

Primers CR-4 and CR-5 from García-Rodríguez et al.<br />

(1998) amplified a 410 bp portion of the mtDNA control<br />

region displacement loop for 113 individuals; 101 live<br />

captures and 12 carcasses. All PCR amplifications were<br />

carried out on a PTC-200 thermal cycler (MJ Research,<br />

Waltham, MA, USA). The PCR reaction conditions followed<br />

Kellogg (2008). Representatives from each haplotype<br />

and any ambiguous sequences were sequenced in both<br />

directions to ensure the accuracy of nucleotide designations.<br />

Microsatellite DNA analysis<br />

A panel of 16 polymorphic microsatellite primers (García-<br />

Rodríguez et al., 2000; Pause et al., 2007) was PCR amplified<br />

using 118 individuals. The study included 88 individuals<br />

from the SLS system (12 Northern Lagoon and 76 Southern<br />

Lagoon), 21 from BCC, two from PL and seven amplifiable<br />

carcass samples. PCR conditions followed Kellogg (2008;<br />

Table 1). MgCl2 concentrations were 3 mM, except for<br />

TmaH13, TmaKb60 and TmaSC5, which required 2 mM.<br />

Fragment analysis was performed on an Applied Biosystems<br />

(Foster City, CA, USA) ABI 3730 Genetic Analyzer. All<br />

individuals amplified at 14 or more loci. The Florida data<br />

were kindly provided for use in this paper (K.P. Tucker<br />

et al., unpubl. data).<br />

Statistical analysis<br />

Mitochondrial DNA analysis<br />

The degree of differentiation, F ST and F ST, betweentheBCC<br />

and SLS and between Belize and Florida was calculated using<br />

ARLEQUIN 3.1 (Excoffier, Laval & Schneider, 2005). Comparisons<br />

with PL were not conducted due to the small sample size.<br />

Estimates of sequence divergence used the Kimura twoparameter<br />

genetic distance model (Kimura, 1980; Jin & Nei,<br />

1990). Lastly, Tajima’s D of selective neutrality, genetic<br />

diversity (h), nucleotide diversity (p), number of polymorphic<br />

sites (S) and number of nucleotide substitutions (NS) were<br />

calculated (Nei, 1987; Tajima, 1993).<br />

Microsatellite DNA analysis<br />

The level of polymorphism was estimated by the observed<br />

(HO) and expected heterozygosity (HE) and the average<br />

Animal Conservation 13 (2010) 592<strong>–</strong>602 c 2010 The Authors. Animal Conservation c 2010 The Zoological Society of London


M. E. Hunter et al.<br />

Table 1 Characteristics and PCR conditions for the 16 microsatellite loci implemented in the Belize manatee samples<br />

Locus name Tm ( 1C) BSA NA HO HE PIC NE TmaA02 56 2 0.310 0.349 0.534 1.537<br />

TmaE1 60 + 4 0.615 0.619 1.121 2.622<br />

TmaE02 55 2 0.265 0.300 0.477 1.429<br />

TmaE7 62 + 4 0.299 0.323 0.627 1.477<br />

TmaE08 58 3 0.701 0.657 1.084 2.9<strong>17</strong><br />

TmaE11 56 5 0.761 0.757 1.488 4.123<br />

TmaE14 62 + 4 0.564 0.573 1.035 2.340<br />

TmaE26 60 4 0.381 0.341 0.622 1.518<br />

TmaF14 58 2 0.390 0.403 0.593 1.675<br />

TmaH13 58 + 2 0.359 0.355 0.540 1.550<br />

TmaJ02 60 2 0.504 0.482 0.675 1.932<br />

TmaK01 54 2 0.530 0.390 0.578 1.638<br />

TmaKb60 56 5 0.416 0.483 0.782 1.935<br />

TmaM79 56 + 3 0.627 0.593 0.974 2.454<br />

TmaSC5 58 3 0.410 0.366 0.663 1.577<br />

TmaSC13 58 2 0.162 0.230 0.391 1.298<br />

Mean 3.1 0.456 0.451 0.762 2.001<br />

The optimized annealing temperature (T m), BSA requirement of 0.4 mg mL 1 , number of alleles (N A), the observed and expected heterozygosity<br />

(H O and H E), polymorphic information content (PIC), and effective number of alleles (N E).<br />

number of alleles per locus (NA) using GenAlEx 6.2 (Peakall<br />

& Smouse, 2006). Departures from linkage disequilibrium<br />

and Hardy<strong>–</strong>Weinberg equilibrium (HWE) were tested (dememorization<br />

10 000, batches 100, iterations per batch<br />

5000) in GENEPOP 4.0 (Raymond & Rousset, 1995) and<br />

Bonferroni’s corrections were applied for multiple comparisons.<br />

To assess overall genetic differentiation at the population<br />

level, GenAlEx 6.2 calculated F ST using the infinite<br />

alleles model and RST using the stepwise mutation model<br />

through an AMOVA. Comparisons included SLS, BCC and<br />

recovered carcasses, and Belize and Florida.<br />

GENECAP (Wilberg & Dreher, 2004) calculated the unbiased<br />

probability of identity (PID), which is the probability that two<br />

individuals drawn at random from a population will have the<br />

same genotype at the assessed loci (Paetkau & Strobeck, 1994)<br />

and P(ID)sib, a related more conservative statistic for calculating<br />

PID among siblings (Evett & Weir, 1998). The program<br />

additionally searched for duplicate genotypes. GenAlEx 6.2<br />

was used to estimate the shadow effect and P(ID)observed to<br />

determine whether unbiased P(ID) or P(ID)sib is the more<br />

accurate probability (Mills et al., 2000; Waits, Luikart &<br />

Taberlet, 2001) following Hunter et al. (2010).<br />

The relationship between genetic and geographic distance<br />

was evaluated to test for isolation by distance within Belize<br />

and for Belize and Florida based on 10 000 randomizations<br />

in GenAlEx 6.2. BOTTLENECK 1.2.02 evaluated heterozygote<br />

excess under the two-phased model using the Wilcoxon signranked<br />

test (Piry, Luikart & Cornuet, 1999). The M-ratio<br />

test used M_P_Val.ex to measure the proportion of unoccupied<br />

allelic states, which is lowered during a population<br />

reduction (Garza & Williamson, 2001). Input values included<br />

a mutation rate of 5 10 4 and y=0.274; and the<br />

recommended values of Dg=3.5 and ps=0.9. Datasets<br />

using seven or more loci and the appropriate mutation<br />

model can be assumed to have experienced a reduction in<br />

Reduced Belize manatee dispersal and genetic variation<br />

population size with an Mo0.68. The coefficient of genetic<br />

relatedness, r xy (Queller & Goodnight, 1989), was used to<br />

test the genetic variability within Belize using 95% confidence<br />

intervals in GenAlEx 6.2. Intra-populational relatedness<br />

should be significantly higher in populations that have<br />

undergone bottlenecks or founding events.<br />

To test the distribution of rxy between all pairs of<br />

individuals, KINGROUP 2 simulated expected unrelated and<br />

full-sib pairwise relatedness (10 000 pairs) from observed<br />

allele frequencies and the rxy values were plotted to compare<br />

the distribution (Konovalov, Manning & Manning, 2004).<br />

A deviation from random expectation could result from<br />

nonrandom mating among related individuals and/or overrepresentation<br />

from a few families, suggesting an excess of<br />

inbred individuals in the population.<br />

The program STRUCTURE 2.3.2 was used to identify the<br />

genetic subdivision within Belize and the genetic relationship<br />

and ancestral source populations of Belize and Florida<br />

manatees (Pritchard, Stephens & Donnelly, 2000). The most<br />

probable number of populations, K, was determined by<br />

evaluating the likelihood of the posterior probability [L(K);<br />

Pritchard, Wen & Falush, 2007] and by calculating DK, an<br />

ad hoc quantity related to the change in posterior probabilities<br />

between runs of different K values (Evanno, Regnaut &<br />

Goudet, 2005) in STRUCTURE HARVESTER (Earl, 2009). Simulations<br />

were conducted using the admixture model of<br />

K=1<strong>–</strong>10 with 10 000 repetitions of MCMC, following the<br />

burn-in period of 10 000 iterations with and without a priori<br />

‘population’ information using the admixture model. The<br />

LOCPRIOR setting was used to detect cryptic structure by<br />

providing priors for the Bayesian assignment process based<br />

on the sample location. The LOCPRIOR model allows structure<br />

to be detected with lower levels of divergence and is not<br />

biased towards detecting structure when it is not present<br />

(Hubisz et al., 2009). Individual assignment success was<br />

Animal Conservation 13 (2010) 592<strong>–</strong>602 c 2010 The Authors. Animal Conservation c 2010 The Zoological Society of London 595


Reduced Belize manatee dispersal and genetic variation M. E. Hunter et al.<br />

recorded as the likelihood of assignment (Q). Individuals<br />

were assigned to a cluster with QZ60%.<br />

Results<br />

Mitochondrial DNA analysis<br />

The Belize samples contained three haplotypes (A03, A04 and<br />

J01). The BCC samples were comprised of the A04 (n=4) and<br />

J01 (n=15) haplotypes. The SLS individuals contained the A04<br />

(n=52), J01 (n=22) and A03 (n=6) haplotypes. A04 (n=1)<br />

and J01 (n=1) were identified in the PL samples. Recovered<br />

carcasses contained the A04 (n=6) and J01 (n=6)haplotypes.<br />

MtDNA sequence divergence estimates were h=0.534 0.025<br />

and p=0.031 0.015. Within Belize, Tajima’s D was large and<br />

positive (D=4.118) consistent with a population bottleneck,<br />

however it was not significant (Po1.000); therefore the<br />

null hypothesis of selective neutrality cannot be rejected.<br />

Genetic differentiation estimates between BCC and SLS<br />

were FST=0.078 (Po0.045 0.015) and FST= 0.036<br />

(Po0.712 0.052). Twenty-eight variable sites and nucleotide<br />

substitutions (6.8%) were identified in the three haplotypes.<br />

Within Florida, one haplotype was observed (A01, p=0.000;<br />

unpubl. data; Garc´ ıa-Rodríguez et al., 1998). Belize and Florida<br />

mtDNA sequence divergence estimates were FST=0.626 (Po0.00001 0.0) and FST=0.302(Po0.00001 0.0).<br />

Microsatellite DNA analysis<br />

The 16 nuclear microsatellite markers had lower levels of<br />

variation [HE=0.453 (0.230<strong>–</strong>0.757); HO=0.456 (0.162<strong>–</strong><br />

0.761), N A=3.1 (2<strong>–</strong>5); Table 1] than the Florida population<br />

over 14 loci (K.P. Tucker et al., unpubl. data). In Belize,<br />

TmaKb60 and TmaSC13 had evidence of null alleles due to a<br />

heterozyogote deficiency. Deviation from HWE was found<br />

for TmaKb60 and TmaK01, even after a sequential Bonferroni<br />

adjustment (Po0.001). After 120 comparisons, linkage<br />

disequilibrium was observed between TmaE1 and TmaE14<br />

(Po0.05). FIS was 0.012 overall, suggesting slight inbreeding<br />

in the population. The error rate was determined by<br />

re-genotyping 11% of the individuals; no errors or duplicate<br />

genotypes were detected.<br />

Genetic differentiation between the BCC, SLS and recovered<br />

carcass samples was significant (Po0.05), although low<br />

(Table 2). The majority of private alleles were detected in<br />

SLS (n=5) as compared with BCC (n=1). The genetic<br />

distances showed a positive relationship with geographical<br />

distances between the BCC and SLS [r=0.20, (Po0.001)]<br />

and Belize and Florida [(r=0.48, (Po0.001)]. Pairwise F ST<br />

and R ST values for Belize and Florida were 0.141 and 0.082<br />

(Po0.001), respectively, supporting the subspecies distinction.<br />

Private alleles were detected for Florida (n=<strong>17</strong>) and<br />

Belize (n=1).<br />

The loci produced an unbiased P(ID) estimate of<br />

4.55E 08 and a P(ID)sib estimate of 3.60E 04, in which<br />

unrelated individuals could be identified in a population<br />

size of 4.7E+03 and 53, respectively, with a shadow<br />

effect of 1.00. Calculations in GenAlEx determined that<br />

P (ID)observed was nearly identical to that of unbiased P (ID),<br />

demonstrating that relatedness in this population is unlikely<br />

to affect match probability, because the population is<br />

o4.7E+03.<br />

The BCC and SLS Mantel test was positive and significant<br />

(Po0.001). All mutation model heterozygosity excess tests<br />

were significant (P 0.05), with low heterozygosity deficiency<br />

to excess ratios, SMM (4:12), IAM (2:14) and TPM (3:13).<br />

However, the mode shift test was L-shaped with no distortion<br />

of allele frequencies. The M-ratio was 0.54, indicative<br />

of a recent reduction in population size. The Belize intrapopulational<br />

pairwise relatedness was rxy=0.184 (95%<br />

CI=0.<strong>17</strong>4<strong>–</strong>0.185), and fell above of the 95% CI ( 0.022<strong>–</strong><br />

0.015) for permutations assuming random mating.<br />

The empirical distribution of pairwise relatedness values<br />

was low (mean=0.002, range=0.852<strong>–</strong>0.815), but not significantly<br />

different from the simulated distribution under<br />

random mating (mean=0.144; Mann<strong>–</strong>Whitney U-test,<br />

Table 2 Genetic differentiation and haplotype proportions of Belize manatees in the Southern Lagoon System (SLS), Belize City Cayes (BCC), and<br />

recovered carcasses (CAR)<br />

Geographic region<br />

(a) Genetic differentiation<br />

SLS BCC CAR<br />

SLS <strong>–</strong> 0.031 0.045<br />

BCC 0.041 <strong>–</strong> 0.025<br />

CAR 0.215 0.033 <strong>–</strong><br />

(b) Haplotype proportions<br />

F ST values (above diagonal) and R ST values (below diagonal) generated from a survey of 16 microsatellite loci. Statistically significant values are in<br />

italics (P 0.01).<br />

Pie charts indicating the haplotype proportions of the A03 (black), A04 (dark gray), and J01 (light gray) in the SLS (n=80), BCC (n=19) and CAR<br />

(n=12) samples.<br />

596<br />

Animal Conservation 13 (2010) 592<strong>–</strong>602 c 2010 The Authors. Animal Conservation c 2010 The Zoological Society of London


M. E. Hunter et al.<br />

P40.05) or the full-sib distribution (Mann<strong>–</strong>Whitney U-test,<br />

P40.05; Fig. 2). However, the full-sib pairwise relatedness<br />

indicated a full-sib detection threshold of 0.21 with 19.9% of<br />

the pairs potentially representing full-sibs under this criterion,<br />

suggesting a strong representation of a few families in<br />

the population.<br />

The Bayesian and resulting Evanno et al. (2005) analyses<br />

strongly supported two genetic clusters (P 1; Fig. 3a and c).<br />

Belize (QAVE=98.5%) and Florida (QAVE=98.6%) had no<br />

admixture and strong assignment of individuals to their<br />

respective populations. When Belize was analyzed alone<br />

without a priori ‘population’ information, one population<br />

was supported (P 1). Alternatively, under the LOCPRIOR<br />

model, K=2 was supported (P 1; Fig. 3b) and a single<br />

genetic cluster or more than two were rejected (P 0, K=1,<br />

3<strong>–</strong>10). DK=2 also had weak support as the largest modal<br />

Figure 2 Distributions of pairwise relatedness values r xy (Queller &<br />

Goodnight, 1989) for the Belize manatee population (gray bars).<br />

Simulated unrelated (solid line) and full-sib (dashed line) curves are<br />

based on 10 000 iterations.<br />

value (Fig. 3d). The two clusters corresponded closely with<br />

the sampled geographic regions, the BCC (95% correct<br />

assignment, QAVE=91.6%) and SLS (86% correct assignment,<br />

QAVE=81.0%).<br />

Discussion<br />

Genetic diversity<br />

Reduced Belize manatee dispersal and genetic variation<br />

Trichechus manatus manatus is believed to have inhabited<br />

the majority of the coastal habitat of Central and South<br />

America historically, although no specific census size estimates<br />

are available (Husar, 1978; IUCN, 2007). A few<br />

thousand manatees were estimated in the 18th century in<br />

Belize (Auil, 1998). Today, the subspecies is severely reduced,<br />

rare, or absent in many areas where manatees were<br />

previously identified as abundant (Husar, 1978). Current<br />

population estimates place the Antillean manatee subspecies<br />

at o2500 mature individuals (IUCN, 2007) and Belize at<br />

approximately 1000 individuals after recovering from a<br />

severe decline (Quintana-Rizzo & Reynolds III, 2007). The<br />

low levels of haplotype diversity, microsatellite heterozygosity<br />

and allelic variation found in the Belize manatee<br />

population are characteristic of small, isolated populations<br />

enduring bottlenecks or long-term persecution (Jamieson,<br />

Wallis & Briskie, 2006).<br />

Belize manatees have lower levels of diversity than other<br />

endangered or bottlenecked populations (Gebremedhin<br />

et al., 2009), including a koala population on French Island,<br />

Australia, which was colonized by three animals (HE=0.48,<br />

N A=3.8, 15 loci; Cristescu et al., 2009), the Wanglang giant<br />

panda (HE=0.68, NA=6.2, 13 loci; He et al., 2008), and the<br />

East African cheetah (HE=0.47, NA=3.55, 82 loci;<br />

Figure 3 Plot of mean and standard deviation of the posterior probabilities, L(K), among STRUCTURE 2.3.3 runs for each value of K (1<strong>–</strong>10) for (a)<br />

Florida and Belize and (b) Belize alone. Plot of DK versus K for (c) Florida and Belize and (d) Belize alone.<br />

Animal Conservation 13 (2010) 592<strong>–</strong>602 c 2010 The Authors. Animal Conservation c 2010 The Zoological Society of London 597


Reduced Belize manatee dispersal and genetic variation M. E. Hunter et al.<br />

Driscoll et al., 2002). Examples of species with lower<br />

diversity than the Belize manatee include the Mediterranean<br />

monk seal (HE=0.41, NA=2.3, 15 loci; Paster et al., 2004)<br />

and the critically endangered Amur leopard (H E=0.37,<br />

NA=2.6, 25 loci; Uphyrkina et al., 2002).<br />

The Belize population was also found to have reduced<br />

variation compared with other marine mammal populations.<br />

Trichechus manatus latirostris currently has greater diversity<br />

and a quickly recovering large population, possibly owing to<br />

a shorter and/or less intense bottleneck period. Hunting of<br />

other marine mammal populations, such as the large cetaceans,<br />

occurred for a similar duration and intensity as for the<br />

Antillean manatee, and also resulted in the commercial<br />

extinction of many populations in the 1800s or 1900s.<br />

Populations with intense hunting and smaller historical<br />

sizes, such as the North Atlantic right whale (NARW), appear<br />

to have lost much of their diversity. The NARW was hunted<br />

for an extended period (11th<strong>–</strong>20th century) and was reduced<br />

from 12 000 individuals to the low hundreds, where it<br />

remains today (Waldick et al., 2002). The NARW genetic<br />

diversity reflects the extended and severe bottleneck the<br />

population has endured (NA=4.1, HE=0.46, 11 polymorphic<br />

loci). Alternatively, the related South Atlantic right<br />

whale population experienced less intense whaling that allowed<br />

the population to remain in the thousands and has<br />

higher genetic diversity (Waldick et al., 2002). Genetic studies<br />

have found large diversity, pre-whaling population estimates<br />

and recovered population sizes for fin (Berube et al., 1998),<br />

minke (Pastene, Goto & Kanda, 2009) and humpbacks whales<br />

(Garrigue et al., 2004), and suggest that the cessation of<br />

hunting in the early 20th century, allowed these populations<br />

to remain above 1000 individuals (Roman & Palumbi, 2003).<br />

Though it is reported as once being abundant in the<br />

Caribbean, T. m. manatus historical population sizes were<br />

not likely as large as some whale populations (Husar, 1977,<br />

1978). Further, a protracted period of severe hunting reduced<br />

the regional populations to o1000 individuals and in<br />

598<br />

Figure 4 Plot of posterior probability of assignment<br />

for individuals (vertical lines) based<br />

on Bayesian analysis of variation at 16 microsatellite<br />

loci. Individuals are grouped by population,<br />

(a) Belize and Florida and (b) the Belize<br />

City Cayes and Southern Lagoon System. Red<br />

indicates genetic cluster 1 and green indicates<br />

genetic cluster 2 in each plot.<br />

some cases to effective extinction (O’Shea & Salisbury, 1991;<br />

UNEP/CEP, 1995). This population contraction could have<br />

reduced the genetic diversity in Belize, which is below even<br />

that of the critically endangered NARW. However, lower<br />

dispersal and immigration capabilities in manatees may<br />

limit the exchange and development of inherent diversity in<br />

the taxa.<br />

The detected heterozygosity excess, mean pairwise values,<br />

and recent reduction in the population size suggests that a<br />

bottleneck/founder event, inbreeding, genetic drift and/or a<br />

reproductive skew has increased the relatedness among<br />

Belize manatees. Because diversity is considered necessary<br />

for adaptation to diseases and environmental changes,<br />

erosion of the currently low variation could negatively affect<br />

the population in the future (Reusch & Wood, 2007).<br />

Populations characterized by low diversity and past inbreeding<br />

are at risk for further losses and those not intensely<br />

monitored can become inbred with little warning (Frankham,<br />

1995; Bijlsma, Bundgaard & Boerema, 2000). Alternatively,<br />

inbreeding accumulating gradually in populations<br />

can provide the opportunity for natural selection to remove<br />

deleterious alleles. To date, there are no indications of<br />

decreased fitness due to inbreeding depression in the Belize<br />

population. However, further reduction of population size<br />

could increase inbreeding above the ‘inbreeding threshold,’<br />

resulting in loss of fitness and risk of extinction (Frankham,<br />

1995). Further monitoring could help to detect inbreeding<br />

depression or declines in population size (Schwartz, Luikart<br />

& Waples, 2007).<br />

The immigration of genetically distinct individuals can<br />

substantially reduce the deleterious effects of inbreeding<br />

(Frankham et al., 2002). However, the enhancement of<br />

diversity through this mechanism is limited since many<br />

manatee populations are small, isolated and movement is<br />

restricted by their dependence on fresh water. Manatees<br />

prefer near-shore habitat, have limited coastal ranges and<br />

infrequently move long distances, highlighting the need for<br />

Animal Conservation 13 (2010) 592<strong>–</strong>602 c 2010 The Authors. Animal Conservation c 2010 The Zoological Society of London


M. E. Hunter et al.<br />

detection and management of genetically isolated populations<br />

(Lefebvre et al., 2001; Vianna et al., 2006; Quintana-<br />

Rizzo & Reynolds III, 2007).<br />

Genetic population structure<br />

Genetic differentiation, private alleles and a strong discrepancy<br />

in haplotype proportions indicate restricted admixture<br />

between the BCC and SLS regions (Table 2). A<br />

significant Mantel test suggests a potential barrier to gene<br />

flow and isolation by distance. The BCC and SLS had<br />

disparate proportions of haplotypes, J01 (79%) and A04<br />

(65%), respectively, and greater mtDNA FST values than<br />

microsatellite FST values, indicating female philopatry and<br />

male-biased gene flow. These data also suggest a larger<br />

number of individuals moving from SLS to BCC. Limiting<br />

human activity in the Bar River and Belize River corridors<br />

might encourage movement and breeding between the SLS<br />

and BCC, leading to greater genetic diversity.<br />

Mitochondrial DNA haplotypes have been analyzed<br />

throughout the manatee range and the A03 haplotype has<br />

been found exclusively in the SLS (presented here; Vianna<br />

et al., 2006), suggesting that A03 females remain isolated,<br />

although further analyses of samples from the region are<br />

needed. Reduced movement has also been observed in manatees<br />

tracked using VHF, GPS or UHF transmitters. The<br />

majority of manatees tagged had strong site fidelity, many<br />

remaining within a 15 mile radius of the tagging location (Auil<br />

et al., 2007). The distinct habitats and lack of movement and<br />

admixture indicate that protection and preservation of the<br />

population as at least two genetic populations would best<br />

protect the discrete diversity in each region.<br />

The increasing incidence of boat strike near Belize City<br />

could quickly erode the BCC population size and diversity<br />

(O’Shea & Salisbury, 1991; Auil, 1998, 2007). The recovered<br />

carcasses from the Belize River mouth or to the south were<br />

not statistically different from the BCC samples. The identified<br />

cause of death in the carcasses was boat strike or<br />

undetermined due to decomposition, corresponding to the<br />

high boat traffic near the mouth of the Belize River.<br />

Phylogeographic relationships<br />

Recently, known Florida manatees have been documented<br />

traveling great distances from Florida, including, Cuba, the<br />

Bahamas and Rhode Island (Reid, 2000; Deutsch et al.,<br />

2003; Alvarez-Alemán, Powell & Beck, 2007). Florida<br />

manatees could potentially travel to Mexico along the Gulf<br />

of Mexico coast or by way of the Yucatán Peninsula. In fact,<br />

Mexico and Florida manatees share the A01 mitochondrial<br />

haplotype, possibly relating the two populations. However,<br />

although gene flow between Florida and Belize through<br />

Mexico is possible, the nuclear and mitochondrial DNA<br />

data do not support admixture, indicating possible barriers<br />

to gene flow (Domning & Hayek, 1986).<br />

Radio-tracking technologies have documented manatees<br />

traveling between Mexico and Belize (Auil et al., 2007;<br />

Morales-Vela et al., 2007). Although these movements have<br />

been tracked and the populations share two haplotypes, each<br />

country also contains a private haplotype (A01 and A03,<br />

respectively; Vianna et al., 2006), suggesting a paucity of<br />

female gene flow and incomplete genetic mixing. The limited<br />

relatedness of the three West Indian populations illustrates the<br />

historically low vagility and genetic exchange of the species.<br />

Future directions<br />

Quantifying the degree of migration and diversity in the<br />

Wider Caribbean using genetic tools is recommended to<br />

clarify the current and future role Belize may play in<br />

maintaining the Antillean subspecies in the region. Additionally,<br />

a more thorough examination of genetic connectedness<br />

among West Indian manatee populations<br />

throughout the range will assist in determining breeding<br />

populations and appropriate units of management for conservation.<br />

Further nuclear studies are anticipated to provide<br />

more detail on the genetic relationships. Identification of<br />

isolated or fragmented populations is important as lack of<br />

gene flow has significant deleterious effects on inbreeding,<br />

fitness and population sustainability (Frankham, 1995;<br />

Frankham et al., 2002). The growth and expansion of the<br />

Belize manatee population could also assist in the repopulation<br />

of small or functionally extinct Antillean manatee<br />

populations throughout the Wider Caribbean region<br />

(O’Shea & Salisbury, 1991; Bonde & Potter, 1995).<br />

Acknowledgments<br />

This work was conducted under the USFWS Wildlife Research<br />

Permit MA79<strong>17</strong>21, issued to the USGS, Sirenia Project.<br />

Funding for this project was provided by the USGS, and<br />

the University of Florida College of Veterinary Medicine<br />

Aquatic Animal Health Program. We would like to thank<br />

the Belize Manatee Project that provided access to genetic<br />

samples, logistical support and permission to collaboratively<br />

work under their Belize Forestry Department scientific collection<br />

and research permits CD/72/02 and CD/60/3/03 with<br />

subsequent amendments issued to N. Auil-Gomez and J.A.<br />

Powell. Additionally, Ginger Clark with the UF ICBR and<br />

Sean McCann provided critical technical assistance and<br />

Martha Sudholt and Susana Caballero provided helpful<br />

editorial recommendations. Any use of trade, product, or firm<br />

names is for descriptive purposes only and does not imply<br />

endorsement by the US Government.<br />

References<br />

Reduced Belize manatee dispersal and genetic variation<br />

Alvarez-Alemán, A., Powell, J.A. & Beck, C.A. (2007). First<br />

report of a Florida manatee documented on the North<br />

Coast of Cuba. Sirenews 47, 9<strong>–</strong>10.<br />

Auil, N.E. (1998). Belize Manatee Recovery Plan. UNDP/<br />

GEF Coastal Zone Management Project. Kingston: Caribbean<br />

Environment Programme.<br />

Animal Conservation 13 (2010) 592<strong>–</strong>602 c 2010 The Authors. Animal Conservation c 2010 The Zoological Society of London 599


Reduced Belize manatee dispersal and genetic variation M. E. Hunter et al.<br />

Auil, N.E. (2004) Abundance and distribution trends of the West<br />

Indian manatee in the coastal zone of Belize: Implications for<br />

conservation. Master’s thesis, Texas A&M University.<br />

Auil, N.E., Powell, J.A., Bonde, R.K., Andrewin, K. &<br />

Galves, J. (2007). Belize Conservation Programme 10 year<br />

summary to Liz Claiborne Art Ortenberg Foundation. New<br />

York: Wildlife Trust.<br />

Beck, C. & Forrester, D.J. 1988. Helminths of the Florida<br />

manatee, Trichechus manatus latirostris, with a discussion<br />

and summary of the parasites of sirenians. J. Parasitol. 74,<br />

628<strong>–</strong>637.<br />

Berube, M., Aguilar, A., Dendanto, D., Larsen, F., Di Sciara,<br />

G.N., Sears, R., Sigurjonsson, J., Urban-R, J. & Palsboll,<br />

P.J. (1998). Population genetic structure of North Atlantic,<br />

Mediterranean Sea and Sea of Cortez fin whales, Balaenoptera<br />

physalus (Linnaeus <strong>17</strong>58): analysis of mitochondrial<br />

and nuclear loci. Mol. Ecol. 7, 585<strong>–</strong>599.<br />

Bijlsma, R., Bundgaard, J. & Boerema, A.C. (2000). Does<br />

inbreeding affect the extinction risk of small populations?<br />

Predictions from Drosophila. J. Evol. Biol. 13, 502<strong>–</strong>514.<br />

Blair, D. 2005. Family Opisthotrematidae Poche, 1926. In<br />

Keys to Trematoda: 401<strong>–</strong>406. Jones, A., Bray, R.A. &<br />

Gibson, D.I. (Eds). London: CABI <strong>International</strong> and The<br />

Natural History Museum.<br />

Bonde, R.K. & Potter, C.W. (1995). Manatee butchering sites<br />

in Port Honduras. Gainesville: USGS, Florida Integrated<br />

Science Center, Sirenia Project.<br />

Cristescu, R., Cahill, V., Sherwin, W.B., Handasyde, K.,<br />

Carlyon, K., Whisson, D., Herbert, C.A., Carlsson, B.L.J.,<br />

Wilton, A.N. & Cooper, D.W. (2009). Inbreeding and<br />

testicular abnormalities in a bottlenecked population of<br />

koalas (Phascolarctos cinereus). Wildl. Res. 36, 299<strong>–</strong>308.<br />

Deutsch, C.J., Reid, J.P., Bonde, R.K., Easton, D.E., Kochman,<br />

H.I. & O’Shea, T.J. (2003). Seasonal movements,<br />

migratory behavior, and site fidelity of West Indian manatees<br />

along the Atlantic Coast of the United States. Wildl.<br />

Monogr. 151, 1<strong>–</strong>77.<br />

Dixon, J.D., Oli, M.K., Wooten, M.C., Eason, T.H.,<br />

McCown, J.W. & Cunningham, M.W. (2007). Genetic<br />

consequences of habitat fragmentation and loss: the case of<br />

the Florida black bear (Ursus americanus floridanus). Conserv.<br />

Genet. 8, 455<strong>–</strong>464.<br />

Domning, D.P. & Hayek, L.C. (1986). Interspecific and<br />

intraspecific morphological variation in manatees (Sirenia:<br />

Trichechus). Mar. Mamm. Sci. 2, 87<strong>–</strong>144.<br />

Driscoll, C.A., Menotti-Raymond, M., Nelson, G., Goldstein,<br />

D. & O’Brien, S.J. (2002). Genomic microsatellites as<br />

evolutionary chronometers: a test in wild cats. Genome Res.<br />

12, 414<strong>–</strong>423.<br />

Earl, D. (2009). Structure Harvester v0.3, Available at http://<br />

users.soe.ucsc.edu/ dearl/software/struct_harvest<br />

(accessed 28 June 2010).<br />

Evanno, G., Regnaut, S. & Goudet, J. (2005). Detecting the<br />

number of clusters of individuals using the software<br />

STRUCTURE: a simulation study. Mol. Ecol. 14,<br />

2611<strong>–</strong>2620.<br />

600<br />

Evett, I.W. & Weir, B.S. (1998). Interpreting DNA evidence:<br />

statistical genetics for forensic scientists. Sunderland: Sinauer<br />

Associates Inc.<br />

Excoffier, L., Laval, G. & Schneider, S. (2005). Arlequin ver<br />

3.0: an integrated software package for population genetics<br />

data analysis. Evol. Bioinform. Online 1, 47<strong>–</strong>50.<br />

Frankham, R. (1995). Inbreeding and extinction <strong>–</strong> a threshold<br />

effect. Conserv. Biol. 9, 792<strong>–</strong>799.<br />

Frankham, R., Ballou, J.D. & Briscoe, D.A. (2002). Introduction<br />

to conservation genetics. Cambridge: Cambridge<br />

University Press.<br />

Garc ıa-Rodr ´ ıguez, ´ A.I., Bowen, B.W., Domning, D.P.,<br />

Mignucci-Giannoni, A.A., Marmontel, M., Montoya-Ospina,<br />

R.A., Morales-Vela, B., Rudin, M., Bonde, R.K. &<br />

McGuire, P.M. (1998). Phylogeography of the West Indian<br />

manatee (Trichechus manatus): how many populations and<br />

how many taxa? Mol. Ecol. 7, 1137<strong>–</strong>1149.<br />

Garc ıa-Rodr ´ ıguez, ´ A.I., Moraga-Amador, D., Farmerie,<br />

W.G., McGuire, P.M. & King, T.L. (2000). Isolation and<br />

characterization of microsatellite DNA markers in the<br />

Florida manatee (Trichechus manatus latirostris) and their<br />

application in selected sirenian species. Mol. Ecol. 9,<br />

2161<strong>–</strong>2163.<br />

Garrigue, C., Dodemont, R., Steel, D. & Baker, S. (2004).<br />

Organismal and ‘gametic’ capture-recapture using microsatellite<br />

genotyping confirm low abundance and reproductive<br />

autonomy of humpback whales on the wintering<br />

grounds of New Caledonia. Mar. Ecol. Prog. Ser. 274,<br />

251<strong>–</strong>262.<br />

Garza, J.C. & Williamson, E.G. (2001). Detection of reduction<br />

in population size using data from microsatellite loci.<br />

Mol. Ecol. 10, 305<strong>–</strong>318.<br />

Gebremedhin, B., Ficetola, G.F., Naderi, S., Rezaei, H.R.,<br />

Maudet, C., Rioux, D., Luikart, G., Flagstad, Ø., Thuiller,<br />

W. & Taberlet, P. (2009). Combining genetic and ecological<br />

data to assess the conservation status of the endangered<br />

Ethiopian walia ibex. Anim. Conserv. 12, 89<strong>–</strong>100.<br />

Gilpin, M.E. & Soul e, ´ M.E. (1986). Minimum viable populations:<br />

processes of extinction. In Conservation biology: the<br />

science of scarcity and diversity: 19<strong>–</strong>34. Soulé, M.E. (Ed.).<br />

Sunderland: Sinauer Associates.<br />

He, W., Lin, L., Shen, F.J., Zhang, W.P., Zhang, Z.H., King,<br />

E. & Yue, B.S. (2008). Genetic diversities of the giant<br />

panda (Ailuropoda melanoleuca) in Wanglang and Baoxing<br />

nature reserves. Conserv. Genet. 9, 1541<strong>–</strong>1546.<br />

Hoelzel, A.R., Halley, J., Obrien, S.J., Campagna, C., Arnbom,<br />

T., Leboeuf, B., Ralls, K. & Dover, G.A. (1993).<br />

Elephant seal genetic variation and the use of simulationmodels<br />

to investigate historical population bottlenecks.<br />

J. Hered. 84, 443<strong>–</strong>449.<br />

Hubisz, M.J., Falush, D., Stephens, M. & Pritchard, J.K.<br />

(2009). Inferring weak population structure with the assistance<br />

of sample group information. Mol. Ecol. Resour. 9,<br />

1322<strong>–</strong>1332.<br />

Hunter, M.K., Broderick, D., Ovenden, J.R., Tucker, K.P.,<br />

Bonde, R.K., McGuire, P.M. & Lanyon, J.M. (2010).<br />

Animal Conservation 13 (2010) 592<strong>–</strong>602 c 2010 The Authors. Animal Conservation c 2010 The Zoological Society of London


M. E. Hunter et al.<br />

Characterization of highly informative cross-species microsatellite<br />

panels for the Australian dugong (Dugong<br />

dugon) and Florida manatee (Trichechus manatus latirostris)<br />

including five novel primers. Mol. Ecol. Resour. 10,<br />

368<strong>–</strong>377.<br />

Husar, S. (1977) The West Indian manatee (Trichechus manatus).<br />

Wildlife Research Report 7, United States Department<br />

of the Interior. Washington, 22 pp.<br />

Husar, S. (1978). Mammal species: Trichechus manatus. Am.<br />

Soc. Mammal. 93, 1<strong>–</strong>5.<br />

IUCN. (2007). 2008 IUCN Red List of Threatened Species.<br />

Gland.<br />

Jamieson, I.G., Wallis, G.P. & Briskie, J.V. (2006). Inbreeding<br />

and endangered species management: is New Zealand<br />

out of step with the rest of the world? Conserv. Biol. 20,<br />

38<strong>–</strong>47.<br />

Jin, L. & Nei, M. (1990). Limitations of the evolutionary<br />

parsimony method of phylogenetic analysis. Mol. Biol.<br />

Evol. 7, 82<strong>–</strong>102.<br />

Kellogg, M.E. (2008) <strong>Sirenian</strong> conservation genetics and Florida<br />

manatee (Trichechus manatus latirostris) cytogenetics.<br />

Doctoral dissertation, University of Florida.<br />

Kimura, M. (1980). A simple method for estimating evolutionary<br />

rate of base substitution through comparative<br />

studies of nucleotide sequences. J. Mol. Evol. 16, 111<strong>–</strong>120.<br />

Konovalov, D.A., Manning, C. & Henshaw, M.T. (2004).<br />

KINGROUP: a program for pedigree relationship reconstruction<br />

and kin group assignments using genetic markers.<br />

Mol. Ecol. Notes 4, 779<strong>–</strong>782.<br />

Lefebvre, L.W., Marmontel, M., Reid, J.P., Rathbun, G.B. &<br />

Domning, D.P. (2001). Status and biogeography of the<br />

West Indian manatee. In Biogeography of the West Indies:<br />

patterns and perspectives. Second edn. 425<strong>–</strong>474. Woods,<br />

C.A. & Sergile, F.E. (Eds). Boca Raton: CRC Press.<br />

McCarthy, T.J. (1986). The gentle giants of Belize. Part II:<br />

distribution of manatees. Belize Audubon Soc. Bull. 18, 1<strong>–</strong>4.<br />

Mills, L., Citta, J., Lair, K., Schwartz, M. & Tallmon, D.<br />

(2000). Estimating animal abundance using noninvasive<br />

DNA sampling: promise and pitfalls. Ecol. Appl. 10,<br />

283<strong>–</strong>294.<br />

Morales-Vela, B., Padilla-Sald ıvar, ´ J.A., Reid, J. & Butler, S.<br />

(2007). First records of long-distance manatee movements<br />

between Mexico and Belize. Sirenews 47, 12<strong>–</strong>14.<br />

Nei, M. (1987). Genetic distance and molecular phylogeny. In<br />

Population genetics and fishery management: 193<strong>–</strong>223. Ryman,<br />

N. & Utter, F. (Eds). Seattle: University of Washington<br />

Press.<br />

O’Shea, T.J. & Salisbury, J.S. (1991). Belize <strong>–</strong> a last stronghold<br />

for the manatees in the Caribbean. Oryx 25, 156<strong>–</strong>164.<br />

Paetkau, D. & Strobeck, C. (1994). Microsatellite analysis of<br />

genetic variation in black bear populations. Mol. Ecol. 3,<br />

489<strong>–</strong>495.<br />

Pastene, L., Goto, M. & Kanda, N. (2009) Genetic analysis on<br />

stock structure in the Antarctic minke whales from the<br />

JARPA research area based on mitochondrial DNA and<br />

microsatellites. <strong>International</strong> Whaling Commission SC/<br />

Reduced Belize manatee dispersal and genetic variation<br />

D06/J9, p. 22. Report presented to the IWC Scientific<br />

Commitee Workshop.<br />

Paster, T., Garza, J., Allen, P., Amos, W. & Aguilar, A.<br />

(2004). Low genetic variability in the highly endangered<br />

Mediterranean monk seal. J. Hered. 95, 291<strong>–</strong>300.<br />

Pause, K.C., Nourisson, C., Clark, A., Kellogg, M.E., Bonde,<br />

R.K. & McGuire, P.M. (2007). Polymorphic microsatellite<br />

DNA markers for the Florida manatee (Trichechus manatus<br />

latirostris). Mol. Ecol. Notes 7, 1073<strong>–</strong>1076.<br />

Peakall, R. & Smouse, P.E. (2006). GENALEX 6: genetic<br />

analysis in Excel. Population genetic software for teaching<br />

and research. Mol. Ecol. Notes 6, 288<strong>–</strong>295.<br />

Piry, S., Luikart, G. & Cornuet, J.M. (1999). BOTTLE-<br />

NECK: a computer program for detecting recent reductions<br />

in the effective population size using allele frequency<br />

data. J. Hered. 90, 502<strong>–</strong>503.<br />

Pritchard, J.K., Stephens, M. & Donnelly, P. (2000). Inference<br />

of population structure using multilocus genotype<br />

data. Genetics 155, 945<strong>–</strong>959.<br />

Pritchard, J.K., Wen, X. & Falush, D. (2007) Documentation<br />

for Structure Software:Version 2.3. Available at http://<br />

pritch.bsd.uchicago.edu/software. html (accessed 28 June<br />

2010).<br />

Queller, D.C. & Goodnight, G.K. (1989). Estimating relatedness<br />

using genetic markers. Evolution 43, 258<strong>–</strong>275.<br />

Quintana-Rizzo, E. & Reynolds, J. III (2007) Regional management<br />

plan for the West Indian manatee (Trichechus<br />

manatus). CEP Technical Report. Caribbean Environment<br />

Programme, United Nations Environment Programme,<br />

Gosier, Guadeloupe, France.<br />

Raymond, M. & Rousset, F. (1995). GENEPOP (version 1.2):<br />

population genetics software for exact tests and ecumenicism.<br />

J. Hered. 86, 248<strong>–</strong>249.<br />

Reid, J.P. (2000). Florida manatee now resident in the<br />

Bahamas. Sire News 33, 7<strong>–</strong>8.<br />

Reusch, T.B.H. & Wood, T.E. (2007). Molecular ecology of<br />

global change. Mol. Ecol. 16, 3973<strong>–</strong>3992.<br />

Roman, J. & Palumbi, S.R. (2003). Whales before whaling in<br />

the North Atlantic. Science 301, 508<strong>–</strong>510.<br />

Schwartz, M.K., Luikart, G. & Waples, R.S. (2007). Genetic<br />

monitoring as a promising tool for conservation and<br />

management. Trends Ecol. Evol. 22, 25<strong>–</strong>33.<br />

Tajima, F. (1993). Measurement of DNA polymorphism. In<br />

Mechanisms of molecular evolution. Introduction to molecular<br />

paleopopulation biology: 37<strong>–</strong>59.Takahata, N. & Clark,<br />

A.G. (Eds). Tokyo/Sunderland: Japan Scientific Societies<br />

Press/Sinauer Associates Inc.<br />

Thornton, C., Shanahan, M. & Williams, J. (2003). From<br />

wetlands to wastelands: impacts of shrimp farming. London:<br />

Environmental Justice Foundation.<br />

UNEP/CEP. (1995) United Nations Environmental Programme<br />

- Caribbean Environment Programme. Technical<br />

Report No. 35. Regional Management Plan for the West<br />

Indian Manatee, Trichechus manatus.<br />

Uphyrkina, O., Miquelle, D., Quigley, H., Driscoll, C. &<br />

O’Brien, S.J. (2002). Conservation genetics of the Far<br />

Animal Conservation 13 (2010) 592<strong>–</strong>602 c 2010 The Authors. Animal Conservation c 2010 The Zoological Society of London 601


Reduced Belize manatee dispersal and genetic variation M. E. Hunter et al.<br />

Eastern leopard (Panthera pardus orientalis). J. Hered. 93,<br />

303<strong>–</strong>311.<br />

Vianna, J.A., Bonde, R.K., Caballero, S., Giraldo, J.P., Lima,<br />

R.P., Clark, A., Marmontel, M., Morales-Vela, B., De<br />

Souza, M.J., Parr, L., Rodriguez-Lopez, M.A., Mignucci-<br />

Giannoni, A.A., Powell, J.A. & Santos, F.R. (2006).<br />

Phylogeography, phylogeny and hybridization in trichechid<br />

sirenians: implications for manatee conservation. Mol.<br />

Ecol. 15, 433<strong>–</strong>447.<br />

Waits, L.P., Luikart, G. & Taberlet, P. (2001). Estimating the<br />

probability of identity among genotypes in natural<br />

602<br />

populations: cautions and guidelines. Mol. Ecol. 10,<br />

249<strong>–</strong>256.<br />

Waldick, R.C., Kraus, S.S., Brown, M. & White, B.N. (2002).<br />

Evaluating the effects of historic bottleneck events:<br />

an assessment of microsatellite variability in the<br />

endangered, North Atlantic right whale. Mol. Ecol. 11,<br />

2241<strong>–</strong>2250.<br />

Wilberg, M.J. & Dreher, B.P. (2004). GENECAP: a program<br />

for analysis of multilocus genotype data for non-invasive<br />

sampling and capture-recapture population estimation.<br />

Mol. Ecol. Notes 4, 783<strong>–</strong>785.<br />

Animal Conservation 13 (2010) 592<strong>–</strong>602 c 2010 The Authors. Animal Conservation c 2010 The Zoological Society of London


Tourism and wildlife habituation: Reduced population fitness<br />

or cessation of impact?<br />

J.E.S. Higham *, E.J. Shelton<br />

Department of Tourism, School of Business, University of Otago, P.O. Box 56, Dunedin, New Zealand<br />

article info<br />

Article history:<br />

Received <strong>17</strong> August 2010<br />

Accepted 15 December 2010<br />

Keywords:<br />

Tourism<br />

Wildlife<br />

Habituation<br />

Stimulus control<br />

Tolerance<br />

Sensitisation<br />

Management<br />

1. Introduction<br />

abstract<br />

The concept of wildlife habituation traditionally has been treated<br />

uncritically within the field of nature-based tourism (Edington &<br />

Edington, 1986; Shelton & Higham, 2007). Typically the concept of<br />

habituation is accepted as a unitary phenomenon that is a negative<br />

possible consequence of tourist interactions with animals in the wild<br />

(Higginbottom, 2004; Newsome et al., 2005; Shackley, 1996). This<br />

global and stable behavioural descriptor, habituation, is an unhelpful<br />

way to formulate most observed lack-of-wildlife-response to<br />

human interactions. A more fine-grained behavioural and temporal<br />

approach to understanding wildlife habituation may be of considerable<br />

relevance to the optimal management of tourist interactions<br />

with wild animals.<br />

This paper attempts two points of departure from established<br />

discourses on wildlife habituation. The first challenges the contrast<br />

between the use of habituation in zoological studies (e.g. to mitigate<br />

researcher impacts), and the treatment of habituation in tourism<br />

management as entirely undesirable (Seddon, Ellenberg, & van<br />

Heezik, in press). Second, it asks if in some instances habituation<br />

* Corresponding author.<br />

E-mail addresses: james.higham@otago.ac.nz, jhigham@business.otago.ac.nz<br />

(J.E.S. Higham), eric.shelton@otago.ac.nz (E.J. Shelton).<br />

0261-5<strong>17</strong>7/$ e see front matter Ó 2011 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.tourman.2010.12.006<br />

Tourism Management 32 (2011) 1290e1298<br />

Contents lists available at ScienceDirect<br />

Tourism Management<br />

journal homepage: www.elsevier.com/locate/tourman<br />

Habituation typically is viewed as a negative consequence of human interactions with wildlife<br />

(Higginbottom, 2004; Newsome, Dowling, & Moore, 2005; Shackley, 1996). While animal habituation<br />

commonly is used in the laboratory and field-based zoology studies, attempts to consider deliberate<br />

habituation specifically in a tourism management context (Shelton, Higham, & Seddon, 2004) has been<br />

received unsympathetically by biological scientists and wildlife managers on the grounds that habituation,<br />

by definition, is undesirable. This paper puts forward the case that the global and stable behavioural<br />

descriptor, habituation, is not the most useful way to formulate most observed lack-of-wildlife-response to<br />

visitor approach and observation. It presents an applied behaviour analysis of wildlife habituation that is<br />

situated within learning theory. This analysis differentiates between avoidance/approach behaviours,<br />

tolerance, habituation and sensitisation. This provides a formulative framework for humanewildlife<br />

interactions, that is then considered specifically in terms of tourism businesses seeking to provide<br />

sustainable visitor interactions with wild animals. A tourism management model derived from this critique<br />

of habituation is presented and discussed.<br />

Ó 2011 Elsevier Ltd. All rights reserved.<br />

may be deployed as a deliberate tourism management strategy in<br />

pursuit of cessation of visitor impacts upon focal animals (Nisbet,<br />

2000). These discussions draw upon three sources; thousands of<br />

unstructured personal observations on the part of the authors<br />

derived from employment as wildlife tour guides in New Zealand;<br />

extensive in situ personal observations of humanewildlife encounters<br />

in New Zealand including New Zealand’s sub-Antarctic Islands,<br />

and Ross Sea Dependency (Antarctica); and key empirical literature<br />

(e.g. Bejder, Samuels, Whitehead, Finn, & Allen, 2009; Ellenberg,<br />

Mattern, Seddon, & Jorquera, 2006; Johns, 1996; Knight, 2009;<br />

Romero & Wikelsi, 2002). The discussions that follow draw upon<br />

person observations unless indicated otherwise through reference<br />

citation.<br />

2. Ecotourism and wildlife encounters<br />

Wildlife viewing, once the domain of dedicated enthusiasts, or<br />

‘specialists’ (Duffus & Dearden, 1990), has moved into the mainstream<br />

of commercial tourism (Knight, 2009). With this has come<br />

a proliferation and diversification of opportunities to encounter<br />

wildlife (Higham, Lusseau, & Hendry, 2008). This course of development<br />

has occurred despite an inescapable tension. Knight (2009:<br />

p. 167) identifies a fundamental contradiction in wildlife viewing in<br />

that “wild animals are generally human-averse; they avoid humans


and respond to human encounters by fleeing and retreating to<br />

cover”. A gradual reduction of such avoidance commonly is labelled<br />

habituation. However humanewildlife encounters represent<br />

a complexity of interaction stimuli that render the unitary term<br />

habituation problematic (Bejder et al., 2009). With this point in<br />

mind, it is surprising that wildlife habituation continues to be<br />

treated uncritically within the context of nature-based tourism and<br />

sustainable tourism management.<br />

Successful commercial wildlife viewing requires that visitors are<br />

concentrated in well-defined locations where interactions with<br />

wild animals are predictable and constant (Whittaker, 1997).<br />

Usually, viewing wildlife takes place where sightings are most<br />

consistent, focal animals can be viewed in abundance or, where<br />

spectacular behaviours may be predictably observed and experienced<br />

(e.g. courtship and socialising behaviours). The critical nature<br />

of these sites, in terms of site ecology and wildlife behaviours, raises<br />

two important points; (1) That the behavioural state of wild animals<br />

varies significantly over time (e.g. across times of day, through<br />

different stages of the breeding cycle and across life course) and (2)<br />

that animal responses to external stimuli (e.g. the presence of<br />

tourists) is likely to vary over time, as influenced by these temporal<br />

determinants.<br />

These points raise the challenge of understanding and<br />

responding to the potential adverse impacts of humanewildlife<br />

interactions, not only for the sake of the focal animals, but also from<br />

an experiential standpoint with the aim of (1) reducing avoidance,<br />

flight or retreat responses and (2) mitigating adverse impacts on<br />

animals that may otherwise discontinue critical, site-specific<br />

behaviours e.g. by instigating site abandonment. Furthermore,<br />

increasingly tourists seek to be assured that their mere presence at<br />

wildlife viewing sites (and their associated behaviours) are not to<br />

the detriment of the animals that they seek to experience first hand,<br />

either in terms of the welfare of individual animals or wider population<br />

fitness (Higham & Lusseau, 2004; Muloin, 1998).<br />

Such concerns on the part of tourists (and site managers) are well<br />

founded. Knight (2009) critiques the dichotomy between wildlife<br />

viewing and wildlife hunting as non-consumptive and consumptive<br />

respectively, highlighting that viewing and hunting wildlife have<br />

some fundamental similarities. First, both engage in locating and<br />

identifying target wild animals, which are generally “wary of human<br />

presence and reluctant to expose themselves to human eyes”<br />

(Knight, 2009: p. 169). Such wariness may be expressed through<br />

anti-predatory adaptations, both physical (e.g. camouflage colouration)<br />

and behavioural (e.g. concealment) to avoid detection. The<br />

directed, intensive and sustained tourist gaze offers a second<br />

parallel with hunting (and predation more generally), which also is<br />

likely to trigger alarm and anti-predatory responses in individual<br />

animals or groups of animals.<br />

Tourist satisfaction is commonly associated with close-up,<br />

unconstrained and prolonged interactions with wild animals, the<br />

experience of critical behaviours (e.g. hunting, feeding, socialising<br />

and courtship) and, in some cases, immediate proximity extending<br />

to touch (e.g. Muloin, 1998). Although there are some exceptions to<br />

this rule (Orams, 2000), the importance of managing ‘humane<br />

wildlife viewing interactions’ (as opposed to ‘non-consumptive<br />

viewing’ e given the consumptive nature of pursuit, intensive gaze<br />

and proximal interaction) is evident. Knight (2009: p. <strong>17</strong>3) proposes<br />

that “there appear to be three main ways in which wild animals can<br />

be made available for human viewing: capture and confinement;<br />

habituation and attraction.”<br />

The first, capture and confinement, brings with it behavioural<br />

diminution associated with wild animals being physically removed<br />

from their natural ecosystem. Also, in many cases, the capture and<br />

confinement of ‘charismatic’ wild animals is illegal. The last, attraction,<br />

through interventions such as provisioning (i.e. feeding) is<br />

J.E.S. Higham, E.J. Shelton / Tourism Management 32 (2011) 1290e1298 1291<br />

equally unpalatable due to its association with manipulated wildlife<br />

distributions, compromised feeding (e.g. predatoreprey) relationships,<br />

increased probability of aggressive behaviour (both towards<br />

animals and people) and consequential diminished behaviours<br />

(Orams, 1995). Attraction of animals through supplemental feeding,<br />

while affording close and enduring viewing opportunities, also<br />

compromises the notion of wild, creating indistinct dividing lines<br />

between animals that are wild and those that are confined, tamed or<br />

domesticated (Knight, 2009). Of the three approaches to making<br />

wild animals viewable, this leaves habituation.<br />

Habituation has been defined as “a decrease in the strength of<br />

a response after repeated presentations of a stimulus that elicits that<br />

response” (Mazur, 2006). As such, habituation typically is viewed as<br />

a negative consequence of human interactions with wildlife due to<br />

the likely consequential reduction of population fitness arising from<br />

reduced danger flight response. However, habituation has been<br />

actively applied by zoological scientists in their field studies (i.e. in<br />

the wild), having been pioneered by primatologists such as Jane<br />

Goodall (Tanzania) working with chimpanzees (Pan troglodytes) and<br />

Dian Fossey (Rwanda) with gorillas (Gorilla gorilla) (Knight, 2009).<br />

Interestingly, Fossey’s habituation of Gorillas made possible the<br />

development of Gorilla tourism as an alternative to Gorilla poaching<br />

(Shackley, 1995). Yet habituation has never been treated critically in<br />

the specific context of sustainable commercial wildlife viewing<br />

(Nisbet, 2000).<br />

The relationship between tourists and wild animals is extremely<br />

complex (Bejder et al., 2009). Duffus and Dearden (1990) contend<br />

that this complexity arises from the interplay of three key components<br />

of the wildlife experience; site users, focal wildlife species<br />

(both individual animals and local animal populations) and the wider<br />

ecology of the viewing site. Given these components, considerations<br />

of habituation in ecotourism can be contemplated only in terms of<br />

site users (e.g. visitor management, guiding practice), the focal<br />

wildlife population (i.e. species-level variables and characteristics of<br />

individual animals) and the wider ecology of the focal species (e.g.<br />

spatio-temporal ecology) (Higham et al., 2008). Duffus and Dearden<br />

(1990) highlight also the temporal dimension in so far as tourist<br />

interactions with wildlife animals vary within diurnal, seasonal and<br />

life course timeframes, a point that has been inadequately addressed<br />

in academic discussions of habituation (Bejder et al., 2009).<br />

Following Duffus and Dearden (1990) then, some species of wild<br />

animals, and some individual animals, demonstrate a well-established<br />

response of inquisitiveness to humans or signs of human<br />

activity. Others are intensely private, remain concealed from<br />

onlookers, and flee readily from human interaction. In terms of<br />

inquisitiveness and active interaction, gulls are a ubiquitous example<br />

but rarely are gulls the focal species of interest to ecotourism operators<br />

and their visitors (Shelton & Higham, 2007). Useful approaches<br />

to the task of analysing the behaviour of species that are of tourist<br />

interest should, however, prove to be of considerable value.<br />

Applied behaviour analysis, based on the notion of stimulus<br />

control and situated within learning theory, is well suited to provide<br />

a formulative framework for those humanewildlife interactions that<br />

are of interest to commercial tourism operators. In this respect, the<br />

term habituation has been described as any situation where wildlife<br />

come to tolerate the presence of humans without any obvious signs<br />

of physiological or behavioural response (Shackley, 1996). Under this<br />

definition, habituation may be considered a mitigation or cessation<br />

of impact. The use of the term habituation in tourism and recreation<br />

has been applied also when animals approach people, or scavenge<br />

for food (Newsome et al., 2005). Park managers in North America<br />

attempt to educate the public not to engage in behaviours that will<br />

merge habituation and approach-for-food, particularly in the case of<br />

deer (The National Park Service, 2006a, 2006b) and bears (Davis,<br />

Wellwood, & Ciarniello, 2002).


1292<br />

To be as useful as possible, habituation must be distinguished<br />

from foraging or increased approach behaviour, even though the<br />

three phenomena regularly occur together (Davis et al., 2002). The<br />

starting point for such an undertaking is a critical behavioural<br />

analysis of wildlife habituation. In response to this need, a critical<br />

discussion of animal responses to human presence in presented, in<br />

order to achieve a more fine-grained understanding of habituation<br />

in all of its complexity. This critique provides a platform from which<br />

then to consider habituation within the context of sustainable<br />

wildlife tourism management.<br />

3. A behavioural formulation of wildlife habituation<br />

A behavioural formulation of wildlife habituation should begin<br />

by acknowledging the poor fit between observable animal behaviour<br />

and internal state (Ellenberg et al., 2006). In other words, the<br />

apparent tolerance of some wildlife species to approach and<br />

observation may not necessarily mean that these wild animals are<br />

not being impacted. Bejder et al. (2009) explain that wildlife tolerance<br />

of human stimuli may arise from various factors including:<br />

(1) Displacement: e.g. less tolerant individual animals may be<br />

displaced, resulting in a bias towards more tolerant animals<br />

that remain at a given site.<br />

(2) Physiology: e.g. reduced responsiveness to human stimuli due<br />

to physiological impairment.<br />

(3) Ecology: e.g. lack of suitable adjacent habitat to which animals<br />

may otherwise relocate.<br />

This poor fit has been demonstrated in the case of penguins<br />

(Ellenberg et al., 2006) where artificial eggs placed under incubating<br />

birds record elevated heart rate in birds that outwardly appeared to<br />

be unaffected when approached (Nimon, Schroter, & Stonehouse,<br />

1995). An understanding of habituation must acknowledge both<br />

the presence and absence of specific behaviours, and that impacts of<br />

significance at both the individual and species level may arise from<br />

both behavioural classes. The presence or absence of behaviours<br />

may vary at different scales of analysis, from individual animals to<br />

Table 1<br />

A critical behavioural formulation of the unitary term habituation.<br />

J.E.S. Higham, E.J. Shelton / Tourism Management 32 (2011) 1290e1298<br />

entire populations (Bejder et al., 2006; Lusseau, 2003; Shelton &<br />

Higham, 2007), and is dynamic over time. Following Bejder et al.<br />

(2009), a behavioural formulation of habituation in the specific<br />

context of wildlife-based tourism may be presented in four parts; 1.<br />

Avoidance/approach behaviours, 2. Tolerance, 3. Habituation and 4.<br />

Sensitisation (Table 1).<br />

3.1. Avoidance/approach behaviour<br />

An important starting point for any critical discussion of habituation<br />

is an understanding of the complexity of avoidance and<br />

approach behaviours. The avoidance behaviours of cetaceans when<br />

engaged in interactions with tourist vessels have been particularly<br />

well researched (Baker, Perry, & Vequist, 1988; Corkeron, 1995;<br />

Lusseau, 2003; Salden, 1988; ). Although responses to boat traffic<br />

vary between species, some show signs of active avoidance, ranging<br />

from altered movement patterns (Bejder, Dawson, & Harraway,1999;<br />

Campagna, Rivarola, Greene, & Tagliorette, 1995; Edds & MacFarlane,<br />

1987; Nowacek, Wells, & Solow, 2001; Salvado, Kleiber, & Dizon,<br />

1992), increases in dive intervals (Baker & Herman, 1989; Baker<br />

et al., 1988; Blane, 1990; Janik & Thompson, 1996; MacGibbon,<br />

1991) and increases in swimming speed (Blane & Jaakson, 1995;<br />

Williams, Trites, & Bain, 2002). Avoidance behaviours in Bottlenose<br />

dolphins (Tursiops truncatus) have been shown to vary at an individual<br />

level (Lusseau, 2003), where male animals show a greater<br />

propensity to avoid boat traffic than females with calves (see Higham<br />

& Lusseau, 2004).<br />

These studies present clear evidence of animal avoidance<br />

behaviours, not only as a result of the presence of boats, but also<br />

arising from the manoeuvring of boats, including sudden changes<br />

in vessel speed and rapid approaches (Constantine, 2001; Gordon,<br />

Leaper, Hartley, & Chappell, 1992; MacGibbon, 1991). In 2006, the<br />

<strong>International</strong> Whaling Commission reached agreement that “there<br />

is compelling evidence that the fitness of individual odontocetes<br />

repeatedly exposed to whale-watching vessel traffic can be<br />

compromised and that this can lead to population-level effects”<br />

(IWC, 2006). This consensus has been reached in light of recent<br />

studies that suggest avoidance responses from cetaceans when<br />

Concept Variables Cases<br />

1. Avoidance/approach Avoidance response (e.g. pre-emptive avoidance) Male Bottlenose dolphins (Tursiops truncatus)<br />

behaviour<br />

Response behaviours (e.g. generalised vigilance, concealment,<br />

discontinuation of critical behaviours)<br />

Male Bottlenose dolphins (Tursiops truncatus)<br />

Stimulus control (i.e. group and individual behaviours) Adele penguins (Pygoscelie adeliae); Little penguins<br />

(Eudyptula minor); Bottlenose dolphins (Tursiops truncatus)<br />

Exploratory approach behaviour Elephant seal (Mirounga leonine)<br />

Opportunistic approach behaviour Buller’s mollymawk (Diomedea bulleri); Giant petrel<br />

(Macronectes giganteus)<br />

Discriminative stimulus Fantail/piwakawaka (Rhipidura fuliginosa); Southern Royal albatross<br />

(Diomedea epomophora)<br />

2. Tolerance Intrinsic tolerance Little penguins (Eudyptula minor); Spotted shags<br />

(Stictocarbo punctatus); Stewart Island shags<br />

(Leucocarbo chalconotus); Little shags (Phalacrocorax melanoleucos)<br />

Selective tolerance New Zealand sea lion (Phocarctos hookeri)<br />

Seasonal tolerance New Zealand wood pigeon/kereru (Hemiphaga novaseelandiae)<br />

3. Habituation Approach habituation New Zealand dabchick (Poliocephalus rufopectus)<br />

Spatio-temporally bound habituation Colonial breeding female New Zealand fur seal (Arctocephalus forsteri).<br />

Socially acquired absence of avoidance behaviour New Zealand fur seal (Arctocephalus forsteri) neonates<br />

Socially acquired habituation Female New Zealand fur seal (Arctocephalus forsteri)<br />

(e.g. new colony recruits)<br />

Reduced physiological arousal Northern Royal albatross (Diomedea epomophora sanfordi)<br />

Reduced avoidance behaviour Yellow-eyed penguin (Megadyptes antipodes)<br />

4. Sensitisation Sensitisation Stoats(Mustela erminea) (i.e. in the laboratory)<br />

Paradoxical sensitisation Galapagos marine iguana (Mblyrhynchus cristatus)


tourist vessels are present, and that those responses disrupt the<br />

behavioural budgets of affected animals. Repeated disturbance can<br />

then lead to displacement from preferred habitat and reduced<br />

fitness at the population level (Bejder et al., 2006) as opposed to<br />

any evidence of habituation.<br />

3.1.1. Stimulus control<br />

The presence or absence of avoidance/approach behaviours<br />

might usefully begin with the treatment of stimulus control. This<br />

term may be used to describe naturally-occurring wild animal<br />

behaviours that are resistant to modification. Adele penguins<br />

(Pygoscelie adeliae) launching from an Antarctic ice floe engage in<br />

grouping and rushing behaviours, presumably to reduce the likelihood<br />

of fatal attack by predators such as Leopard seals (Hydrurga<br />

leptonyx). The same stimulus control applies to Little penguins<br />

(Eudyptula minor) at tourist sites such as the Philip Island Penguin<br />

Parade (Victoria, Australia) and Oamaru Blue Penguin Colony (New<br />

Zealand). In this instance proximity to the landing site and<br />

grouping induces a rushing response which is unchanged despite<br />

millions of unchallenged landings over many generations of birds<br />

(Houston & Russell, 2000). This failure to habituate to coastline<br />

proximity has no obvious evolutionary cost to the species. The<br />

lighting, public address systems and grandstands full of visitors at<br />

these sites do not deter the birds from landing where they have, in<br />

recent history, always landed. This lack of avoidance response has<br />

not changed over time and is uniquely characteristic of this species<br />

of penguin which does not demonstrate habituation of approach or<br />

avoidance behaviour.<br />

Another relevant case involves cetaceans. Lusseau’s (2003)<br />

study of Bottlenose dolphins (Tursiops truncatus) in Doubtful<br />

Sound (Fiordland, New Zealand) found that these marine mammals<br />

entertained visitors on tour boats with their bowriding and leaping<br />

out of the water, such that it might be assumed that dolphins<br />

engage in this behaviour by choice. Power boats, though, act as<br />

a stimulus for dolphins not to rest; that is, bowriding and leaping<br />

are more likely to be under stimulus control and therefore not the<br />

result of choice. There are clear potential negative impacts of<br />

decreased resting on this species, including a disrupted energy<br />

budget and likely compromised reproductive success (Lusseau &<br />

Higham, 2004).<br />

Repeated exposure to tourist vessels does not seem to reduce<br />

the strength of the dolphins’ approach responses and this situation<br />

clearly demonstrates the difficulties inherent in referring to stable<br />

or increasing approach behaviour as habituation. Lusseau and<br />

Higham (2004) suggest the spatial segregation of non-permitted<br />

boats and Bottlenose dolphins through multi-level zoning. An<br />

additional management strategy could, in this case, include<br />

inducing habituation via stimulus mitigation; that is, reducing the<br />

dolphin’s responsiveness to the presence of tourist vessels through<br />

the close management of key stimuli (e.g. boat speed and noise) so<br />

as to counter the stimulus control that these animals appear to act<br />

under.<br />

3.1.2. Exploratory approach behaviour<br />

A further manifestation of approach/avoidance behaviour, with<br />

an apparent temporal (i.e. life course) dimension may be termed<br />

exploratory approach behaviour. Elephant seals (Mirounga leonine)<br />

on sub-Antarctic Campbell Island may be exposed fleetingly to two<br />

or three groups of visitors per year at North West Bay (and not at all<br />

elsewhere on Campbell Island). Juvenile Elephant seals will<br />

approach and investigate tour parties at close range. The lack of<br />

similar interest on the part of adult Elephant seals very probably<br />

reflects a decreasing exploratory behaviour as a consequence of<br />

maturation rather than through habituation (i.e. repeated exposure<br />

to humans) where tourist encounters are so uncommon.<br />

J.E.S. Higham, E.J. Shelton / Tourism Management 32 (2011) 1290e1298 1293<br />

This change in responsiveness raises important questions<br />

around the naturally-occurring contingencies and physiological<br />

processes that operate to reduce exploratory behaviour during<br />

maturation from neonate through juvenile to adult. Repeated nonconsequential<br />

exposure to a range of human behaviours should be<br />

situated somewhere in this process. On mainland New Zealand,<br />

where tourist encounters are more common, adult Elephant seals<br />

will rest ashore, selecting where possible confined and inaccessible<br />

resting sites, and demonstrating complete indifference to approach,<br />

based on a confidence of remaining unmolested. This indifference<br />

may be understood as reduced exploratory approach behaviour<br />

rather than habituation.<br />

3.1.3. Opportunistic approach behaviour<br />

The term opportunistic approach behaviour may be used to<br />

address approach stimuli linked to an individual’s energy budget.<br />

Offshore boat-based tourist encounters with some seabird species<br />

such as Buller’s mollymawk (Diomedea cauta eremita) may result in<br />

close but fleeting observations of birds derived from what may be<br />

descried as ‘investigatory fly past’. Opportunistic approach behaviour<br />

may be a generalisation from the birds’ habit of following fishing<br />

boats whose crews routinely engage in gutting offshore. Giant<br />

petrels (Macronectes giganteus) also will engage in this behaviour.<br />

Distinct from provisioning and scavenging, these birds approach in<br />

response to a specific stimulus. Despite daily exposure to the tourist<br />

boat, and never having been fed, this exploratory approach behaviour<br />

persists over time. Due no doubt to a failure to distinguish<br />

between tourist vessels and fishing boats, and with so little energy<br />

required to complete an investigation, these birds, as an optimal<br />

feeding strategy, casually and routinely approach every boat.<br />

3.1.4. Discriminative stimuli<br />

Related to this, approach behaviours may be further differentiated<br />

from habituation via an understanding of discriminative<br />

stimuli. The behaviour of fantails/piwakawaka (Rhipidura fuliginosa)<br />

is typically misleading. Visitors often encounter the situation where<br />

they are walking along a bush track and suddenly numerous<br />

fantails appear and flutter tantalisingly close-by. Many may incorrectly<br />

assume that it is their presence that has attracted the birds<br />

which are attracted to human presence. Rather, the birds are<br />

attracted to small flies that have risen from the leaf litter as a result<br />

of human disturbance. The attraction to visitors in the bush is<br />

indirect. People in this case act as a discriminative stimulus for<br />

a likely disturbed forest floor, and consequently available prey.<br />

Similarly, Southern Royal albatrosses treat private boats as<br />

discriminative stimuli for fish being scaled and gutted and land in<br />

their wakes. These examples only serve to further confirm the poor<br />

fit between observable behaviour and internal state in terms of<br />

understanding animal behaviours in the wild.<br />

3.2. Tolerance<br />

3.2.1. Intrinsic tolerance<br />

Within investigations of humanewildlife interactions the<br />

apparent tolerance of wild animals to human approach has<br />

remained almost entirely undifferentiated from habituation. The<br />

exception is Bejder et al. (2009: p. 181) who define tolerance as the<br />

“intensity of disturbance that an individual. tolerates without<br />

responding in a defined way”. Tolerance may readily be confused<br />

with the complete absence of tourist impact. Indeed Higham<br />

(1998:529), in reference to viewing Northern Royal albatrosses<br />

(Diomedea epomophora sanfordi), warned that “tolerance should not<br />

disguise the fact that serious impacts may still take place”. The<br />

manifestations of disturbance in the case of these outwardly<br />

tolerant birds have been revealed only by long-term monitoring and


1294<br />

analysis (Higham et al., 2008). A more critical treatment of tolerance<br />

allows, in the first instance, for consideration of intrinsic tolerance,<br />

which varies both between individual animals and species.<br />

Many species of seabirds will demonstrate intrinsic tolerance to<br />

boat approach, seemingly oblivious to engine vibration and noise<br />

(both mechanical and human). Little penguins (Eudyptula minor)<br />

demonstrate intrinsic tolerance of boat engine vibration and noise,<br />

which indicates that they clearly are able to ignore human presence<br />

across settings (i.e. coastal landing zones and open water). Various<br />

species of New Zealand shag (cormorant) also demonstrate<br />

intrinsic tolerance to human approach (e.g. Spotted shags (Stictocarbo<br />

punctatus), Stewart Island shags (Leucocarbo chalconotus) and<br />

Little shags (Phalacrocorax melanoleucos)) (Shelton & Higham,<br />

2007). However tolerance inevitably is behaviour (stimulus)<br />

specific. Supporting Knight’s (2009) treatment of the tourist gaze is<br />

the fact that an intense and prolonged stare is sufficient to cause<br />

some individual birds (e.g. Spotted shags) to abandon their perches.<br />

3.2.2. Selective tolerance<br />

Tolerance, like habituation, has a spatio-temporal component.<br />

Solitary adult male New Zealand sea lions (Phocarctos hookeri)<br />

demonstrate various responses to human approach which may be<br />

termed selective tolerance (or intolerance). Some adult male New<br />

Zealand sea lions will, when in the water in the autumn months,<br />

respond intolerantly (e.g. aggressive approach, charging and biting)<br />

to the passing kayaks of a commercial operator. Repeated exposure<br />

to kayakers, at this time of year, does not decrease the likelihood of<br />

aggressive behaviour. At other times of the day when hauled out on<br />

a nearby beach the same individual animals will tolerate close<br />

approaches by visitors (Shelton & Higham, 2007). The sharp<br />

contrasts between agonistic and indifferent responses (which<br />

clearly vary on a spatio-diurnal basis) are almost certainly linked to<br />

specific spatio-ecological (i.e. territorial dominance) and temporal<br />

(i.e. reproduction) elements.<br />

3.2.3. Seasonal tolerance<br />

A variation of selective tolerance is seasonal tolerance. The New<br />

Zealand wood pigeon/kereru (Hemiphaga novaseelandiae) frequents<br />

some urban areas in New Zealand and can appear to be indifferent<br />

to everyday human activity. Closer analysis reveals seasonal variations<br />

in the species’ tolerance of human stimuli depending on food<br />

availability (Shelton & Higham, 2007). As the season progresses,<br />

and food becomes more abundant, kereru increasingly favour<br />

branches higher up in trees, and seasonal responses to human<br />

activity appear to change.<br />

In this case, it is clearly necessary to elucidate precisely what<br />

role people are playing as discriminative stimuli. People minding<br />

their own business, presumably measured by their unmodified<br />

patterns of movement, uninterrupted chatter and non-birddirected<br />

gaze, act as discriminative stimuli for approach tolerance.<br />

A sudden silence, ceasing walking, or bird-directed gaze act as<br />

stimuli for flight. The paradox for the commercial wildlife tourism<br />

operator who intends that visitors experience pigeons close-up is<br />

that the best advice to be given is to ignore them. Natural settings<br />

do not offer the frequency of close contact that would be required<br />

to attempt deliberately to manipulate the birds’ existing responses.<br />

These two examples illustrate a more general concern with how<br />

humans can best interact with multiple species of birds in urban<br />

areas (Blumstein, Fernandez-Juricic, Zollner, & Garity, 2005).<br />

3.3. Habituation<br />

3.3.1. Approach behaviour<br />

Bejder et al. (2009: p. 181) define habituation as the “relative<br />

persistent waning of a response as a result of repeated stimulation<br />

J.E.S. Higham, E.J. Shelton / Tourism Management 32 (2011) 1290e1298<br />

which is not followed by any kind of reinforcement”. Shackley<br />

(1996) observes that wildlife may habituate to the presence of<br />

humans, showing no obvious signs of physiological or behavioural<br />

response. If so, such a lack of response may fruitfully be explored<br />

within the context of sustainable wildlife tourism and cessation of<br />

impact. However, habituation can be assumed to have important<br />

spatio-temporal dimensions. New Zealand dabchick (Poliocephalus<br />

rufopectus) may become habituated to human approach following<br />

multiple exposures (Bright, Reynolds, Innes, & Waas, 2003). Similarly<br />

diurnal or seasonal habituation may take place in some species<br />

of wild animals under certain circumstances. Non-breeding male<br />

New Zealand fur seals (Arctocephalus forsteri) may be placid on<br />

land, where they demonstrate a decreased gross behaviour<br />

response to human approach. The same animals may be aggressive<br />

to swimmers in the water, which acts against describing any individual<br />

animals as being habituated. Nearby breeding colonies of<br />

female New Zealand fur seals that are exposed to regular stimuli<br />

may become unresponsive to approach.<br />

3.3.2. Socially acquired habituation<br />

Furthermore, there is evidence to suggest that the absence of<br />

response behaviours may be socially acquired or learnt (Bejder<br />

et al., 2009) by wild animals. Early exposure, in the presence of<br />

an unresponsive mother, may result in seal pups that have never<br />

perceived tourist approach as a stimulus for flight. Seal pups do<br />

vary in their post-weaning exploratory audacity and some, having<br />

moved around the site once in the presence of humans with no illeffects,<br />

begin to include indifference in their behavioural repertoire.<br />

Each uneventful visit further reduces the discriminative value of<br />

humans. Those adult females who had been born at the colony<br />

could conceivably retain their neonate indifference but those<br />

recruited from other colonies, where avoidance response to less<br />

frequent human approach takes place, may demonstrate socially<br />

acquired habituation through their peers modelling indifference.<br />

Such complexities of seal habituation to human approach,<br />

including the confounding effect of probable previous exposure,<br />

has been investigated by van Polanen Petel (2005) who suggests<br />

a species-specific approach is required in order to manage<br />

humaneseal interactions optimally.<br />

3.3.3. Reduced physiological arousal<br />

Habituation in the form of complete indifference may be<br />

usefully differentiated from reduced physiological arousal, in which<br />

behavioural responses remain evident, but with reduced energy<br />

expenditure and arousal. At Taiaroa Head (Dunedin, New Zealand)<br />

Northern Royal albatross (Diomedea epomophora sanfordi) are<br />

closely monitored (Higham, 1998) and tourists are confined to<br />

a soundproof hide glazed with reflective glass. Throughout the<br />

breeding season all individual birds in the colony are exposed to<br />

close contact and attention from conservation staff, including<br />

weekly handling during weighing and measuring. These birds do<br />

not abandon their nests but do continue the defencive use of their<br />

beaks, as do Northern Royal albatrosses on the remote Campbell<br />

Island. Despite having been handled periodically over the years<br />

since their hatching these birds did not seem to have habituated to<br />

this particular human behaviour. However, although there had<br />

been no apparent change in the gross motor component of the<br />

birds’ defence rituals, there may well have been reductions in<br />

human-induced physiological arousal. Thus, the defence behaviour<br />

persists, but with reduced determination or persistence.<br />

3.3.4. Reduced avoidance behaviour<br />

A fourth manifestation of habituation may be termed reduced<br />

avoidance behaviour. This arises in cases where approach/avoidance<br />

is at the discretion of individual animals. While it is conceivable


that approach behaviour may be habituated the more likely case in<br />

wild animal populations is that avoidance behaviour may decrease<br />

over time. This appears to be the case in Yellow-eyed penguin<br />

(Megadyptes antipodes) colonies that are exposed to increasing<br />

tourist interactions, where established members of the colony may<br />

become less timid over time, but there may be a simultaneous<br />

reduction in the recruitment of prospecting birds. The impact of<br />

humans on this most phlegmatic but inquisitive of penguins is the<br />

subject of ongoing study (Seddon, Smith, Dunlop, & Mathieu, 2004).<br />

This work is predicated on the fact that measuring penguin habituation<br />

using only approach/avoidance behaviour is inappropriate<br />

since physiological arousal fluctuates independently of observable<br />

behaviour (Ellenberg & Mattern, 2004) and the relationship varies<br />

both between individual penguins of the same species and different<br />

penguin species (Ellenberg et al., 2006).<br />

3.4. Sensitisation<br />

Finally, it is important to attempt to differentiate between<br />

sensitisation and habituation. Usually, sensitisation refers to an<br />

increase in the strength of a response upon repeated or ongoing<br />

exposure to a stimulus that has significant consequences (Bejder<br />

et al., 2009). However, it is noteworthy that repeated exposure to<br />

tourists appeared to lower the production of corticosterone,<br />

commonly a stress hormone, in Galapagos marine iguana (Romero &<br />

Wikelsi, 2002). Lower corticosterone production in response to<br />

human interaction, compared with animals who have no human<br />

interaction, seems to demonstrate sensitisation, the opposite of<br />

habituation, but in this case the lowering of corticosterone production,<br />

although demonstrating sensitisation, may legitimately be<br />

labelled paradoxical sensitisation. Of course one study is inconclusive<br />

J.E.S. Higham, E.J. Shelton / Tourism Management 32 (2011) 1290e1298 1295<br />

but it does raise an intriguing possibility. If robust, this finding<br />

implies that chronic sub-optimal physiological processes may be<br />

a naturally-occurring phenomenon in some species of wild animals.<br />

4. Wildlife habituation and sustainable nature-based<br />

tourism: a management model<br />

This discussion addresses the proposition that ecotourism<br />

operators who are committed to providing a sustainable naturebased<br />

product will be interested to know how their activities may<br />

be influencing the environments in which they do business. The<br />

inherent tensions between business practice and sustainability in<br />

nature-based tourism have been well described (Fennell, 2003;<br />

McKercher, 1993a, 1993b, 1998; Newsome et al., 2005). The<br />

preceding discussions allow for the development of a management<br />

model for sustainable wildlife tourism interactions (Fig. 1) based on<br />

a critical behavioural formulation of habituation.<br />

Following Duffus and Dearden (1990) Fig. 1 is constituted<br />

principally by three key elements; site users, focal animals (individuals<br />

or local populations) and the wider ecology of the wild<br />

animals engaged with human approach and observation. Such<br />

a representation recognises that any humanewildlife interactions<br />

are spatially bound; that is, human interactions with individual<br />

wild animals can take place in a variety of settings that vary across<br />

a range of habitats where different behaviours (e.g. migration,<br />

feeding, resting, socialising, breeding, raising neonates etc.) take<br />

place. Animal responses to human approach and interaction inevitably<br />

vary between settings.<br />

The model recognises also the temporal variation that exists in<br />

terms both of site users (e.g. timing and duration of visit, and<br />

seasonal context), and focal animals (e.g. diurnal, seasonal and life<br />

Fig. 1. Management model for sustainable wildlife tourism interactions based on a critical behavioural formulation of the term habituation.


1296<br />

course), which influence the outcomes of humaneanimal interactions.<br />

Avoidance/approach behaviours, tolerance, habituation and<br />

sensitisation can fruitfully be considered within this context.<br />

A critical understanding of spatio-temporal ecology and the interactions<br />

of tourists and focal animals can then be applied in terms of<br />

tourism management planning and consideration of conscious<br />

management interventions (Fig. 1). This would require a critical<br />

understanding of the unitary term habituation, which is attempted<br />

in this article. It would also necessitate a mixed (biological and<br />

social) science platform of research to inform an integrated and<br />

adaptive management approach (Higham, Bejder, & Lusseau, 2009).<br />

5. Discussion<br />

As nature-based products are offered in more and more remote<br />

places (i.e. previously unexploited ecosystems) it becomes<br />

increasingly likely that any given individual animal will encounter<br />

humans, engaged in particular behaviours, at some stage of its life.<br />

The animal may approach the human, as neonates and juveniles of<br />

many species do. If, on repeated exposure to these particular<br />

human behaviours, this approach behaviour decreases, and is not<br />

replaced by escape/avoidance behaviour, then the animal may<br />

accurately be described as being habituated to the human behaviours<br />

to which it has been exposed, in a particular spatio-temporal<br />

setting.<br />

Any habituation/sensitisation process is dynamic, and may<br />

reflect the hormonal state of the individual animal at the time of<br />

exposure to the particular human behaviour. If the individual<br />

animal demonstrated indifference to the human behaviour right<br />

from the start then habituation is not the best term to use to<br />

describe the situation; tolerance or resilience may be more useful.<br />

Similarly, animals may respond to human activity through<br />

increases in physiological processes, commonly grouped under the<br />

term arousal. Decreasing physiological arousal upon repeated<br />

exposure indicates habituation, increased arousal indicates sensitisation.<br />

The unusual situation where an initially high level of<br />

arousal may be reduced by exposure to human activity may be<br />

tentatively labelled paradoxical sensitisation. It is possible also for<br />

wildlife to decrease one response to human presence, for example<br />

fleeing, but to increase another response. Chimpanzees’ activity<br />

rates in Kibale Forest, Uganda, habituate to the presence of<br />

increasing numbers of observers but vocalisation increases (Johns,<br />

1996).<br />

It is useful at this point to address two issues. Firstly, why is it<br />

important to specify to what human behaviours, in what settings<br />

and at what times do individual animals habituate, and to avoid the<br />

simpler descriptor; habituation to humans? Habituation to<br />

humans, by definition, involves decreased vigilance in response to<br />

human presence, decreased communication of alarm upon coming<br />

across humans and/or decreased avoidance of, or fleeing from,<br />

humans. The argument then is that habituated animals are made<br />

vulnerable in that habituation to one potential predator, Homo<br />

sapiens will induce a generalised reduction of predator-preparedness<br />

(Beale & Monaghan, 2004).<br />

We argue that this concern evaporates when habituation is<br />

more usefully formulated. If habituation is related to spatially and<br />

temporally constrained specific human behaviours, rather than<br />

simply to human presence, then it is reasonable to hypothesise that<br />

individual animals will retain predator-preparedness outside of<br />

these parameters. Discriminating between human behaviours is<br />

consistent with laboratory-based enquiry that has demonstrated<br />

that many species have the ability to discriminate even between<br />

individual human beings performing the same behaviours (Davis,<br />

2002). Predator-preparedness training in birds, in common with<br />

exploratory behaviours (Huber, Rechberger, & Taborsky, 2001), can<br />

J.E.S. Higham, E.J. Shelton / Tourism Management 32 (2011) 1290e1298<br />

occur through social learning. For example adult Sandhill cranes<br />

teach predator-preparedness by modelling vigilance and predatorresponse<br />

even to unrelated captive-reared conspecific chicks<br />

(Heatley, 1995). There is debate about how possible habituation to<br />

human activity acquired while in captivity, for example in a rehabilitation<br />

facility, should be managed. A fine-grained analysis is<br />

available addressing the issue of how much deliberate training-forrelease<br />

should take place, what such training should comprise, and<br />

exactly how humans should be involved in its delivery (Bauer,<br />

2005).<br />

It is reasonable to suppose that habituation can be socially learnt<br />

also in natural settings, and be passed on in a sophisticated rather<br />

than generalised way. Different species teach their offspring<br />

different sensitivities. For cliff-dwelling birds visited by tourists,<br />

size-of-party has been reported to have species-specific effects<br />

(Beale & Monaghan, 2005). It is not unreasonable to suggest that<br />

certain species may teach their young not to respond to a particular<br />

wildlife guide, acting within certain behavioural limits. In such<br />

cases it is critical that individual animals are the unit of analysis in<br />

rigorous longitudinal studies (Bejder et al., 2009).<br />

The second issue that needs to be addressed is one of<br />

management. If, as suggested here, some or all of the behaviours of<br />

many wild species can be brought under stimulus control, then<br />

what happens to the wild in wildlife? The application of stimulus<br />

control techniques to the management of human behaviour is<br />

widely reported in the Applied Behaviour Analysis literature. Using<br />

the nuanced approach to habituation outlined here it is conceivable<br />

that wildlife managers and visitor managers could co-operate to<br />

apply behavioural technologies to structure almost completely the<br />

nature of humanewildlife interactions site-by-site, species-byspecies,<br />

and in some cases to the level of individual animals, taking<br />

into account time of day and time of year. This level of management,<br />

although common in zoos and other captive settings, if<br />

applied to natural settings would challenge directly the motivation<br />

for visiting such sites in the first place. Nonetheless, in certain cases<br />

the conservation outcomes available using such an approach may<br />

be justified.<br />

6. Conclusion<br />

This article highlights the wide variability of wild animal<br />

responses to human approach and interaction. In doing so, it begs<br />

the question; which individual animals or groups of animals<br />

respond in what fashion to what kind of tourist behaviour in what<br />

contexts? Clearly individual (or groups of) animals that respond<br />

passively in one setting, may in other settings not be so sanguine.<br />

For example, where a particular set of contingencies operate to<br />

induce decreasing gross behavioural responsiveness, on land, to<br />

a fairly narrow set of visitor behaviours, in a group of seals who<br />

shared non-breeding status, evidence of spatially and temporally<br />

bounded habituation may exist. The responsiveness of physiological<br />

processes within individual seals, the role of developmental<br />

maturation and the influence of vicarious or social learning through<br />

the observation of calm conspecifics remain largely unexamined,<br />

although available for formulation.<br />

Clearly, a fine-grained behavioural approach is required to<br />

overcome the poor fit between observable wildlife behaviour in the<br />

presence of humans and individuals’ internal states, and to<br />

accommodate the fact that habituation is spatially and temporally<br />

bounded. Given that such a poor fit is unhelpful to individual,<br />

species or whole-of-ecosystem management, it seems desirable<br />

that ecotourism operators actively address likely habituation or<br />

other conditioning to humans as part of their business plan<br />

or concession application. As such, wildlife managers and visitor<br />

managers could co-operate to apply behavioural technologies


to structure almost completely the nature of humanewildlife<br />

interactions. A tourism management model that accommodates<br />

a nuanced understanding of wildlife habituation is presented,<br />

building upon a formulative framework for humanewildlife interactions.<br />

The development of empirical case studies to inform and<br />

monitor tour operator best practices is a valuable avenue of future<br />

research.<br />

Knight (2009) presents three approaches to making wildlife<br />

viewable; capture and confinement, habituation, and attraction.<br />

While of the three, habituation is the least likely to be associated<br />

with diminished behaviours, the question of diminished wildness<br />

remains unanswered. Exploring this question will, no doubt, reveal<br />

a trade off between using habituation as a tourism management<br />

tool and the potential evolutionary costs to focal animal populations,<br />

not only in terms of physiological and behavioural<br />

responses, but also in terms of exposure to disease (Bejder et al.,<br />

2009). One provocative approach, integrated with the field of<br />

neurobiology, will be designer wild species, specifically habituated<br />

for viewing. Alternatively, Knight (2009) observes that learnt<br />

animal behaviours can be unlearnt, giving rise to the notion of the<br />

reversal of habituation. AsKnight (2009: p. 180) comments, “In an<br />

age when our visual appetite for wildlife has never been greater,<br />

there may be good reasons for keeping this appetite in check and<br />

acting to make viewable wild animals unviewable again”. Before<br />

further contemplating the reversal of habituation, a more critical<br />

understanding of habituation itself is required. This amplifies the<br />

need for “studies (that) adopt a long-term experimental design<br />

involving sequential sampling of the same individuals at different<br />

levels of exposure to a disturbance” (Bejder et al., 2009: p. 181). In<br />

the meantime, these thoughts pose intriguing questions that are<br />

available for scholars to debate and research.<br />

References<br />

Baker, C. S., & Herman, L. M. (1989). Behavioral responses of summering humpback<br />

whales to vessel traffic: Experimental and opportunistic observations. Final Report<br />

to the National Park Service. Anchorage, Alaska: Alaska Regional Office.<br />

Baker, C. S., Perry, A., & Vequist, G. (1988). Humpback whales of Glacier Bay, Alaska.<br />

Whalewatcher Fall, 13e<strong>17</strong>.<br />

Bauer, G. (2005). Research training for releasable animals. Conservation Biology,<br />

19(6), <strong>17</strong>79e<strong>17</strong>89.<br />

Beale, C., & Monaghan, P. (2004). Human disturbance: people as predation-free<br />

predators? Journal of Applied Ecology, 41, 335e343.<br />

Beale, C., & Monaghan, P. (2005). Modeling the effects of limiting number of visitors<br />

on failure rates of seabird nests. Conservation Biology, 19(6), 2015e2019.<br />

Bejder, L., Dawson, S. M., & Harraway, J. A. (1999). Responses by Hector’s dolphins to<br />

boats and swimmers in Porpoise Bay, New Zealand. Marine Mammal Science,<br />

15(3), 738e750.<br />

Bejder, L., Samuels, A., Whitehead, H., Finn, H., & Allen, S. (2009). Impact assessment<br />

research: use and misuse of habituation, sensitisation and tolerance in<br />

describing wildlife responses to anthropogenic stimuli. Marine Ecology Progress<br />

Series, 395, <strong>17</strong>7e185.<br />

Bejder, L., Samuels, A., Whitehead, H., Gales, N., Mann, J., Conner, R., et al. (2006).<br />

Decline in relative abundance of bottlenose dolphins exposed to long-term<br />

disturbance. Conservation Biology, 20(6), <strong>17</strong>91e<strong>17</strong>98.<br />

Blane, J. M. (1990). Avoidance and interactive behaviour of the Saint Lawrence<br />

beluga whale (Delphinapterus leucas) in response to recreational boating. M.A.<br />

thesis, Canada: University of Toronto.<br />

Blane, J. M., & Jaakson, R. (1995). The impact of ecotourism boats on the Saint<br />

Lawrence beluga whales. Environmental Conservation, 21(3), 267e269.<br />

Blumstein, D., Fernandez-Juricic, E., Zollner, P., & Garity, S. (2005). Inter-specific<br />

variation in avian responses to human disturbance. Journal of Applied Ecology,<br />

42(5), 943e953.<br />

Bright, A., Reynolds, G., Innes, J., & Waas, J. (2003). Effects of motorised boat passes<br />

on the time budget of New Zealand dabchick, Poliocephalus rufopectus. Wildlife<br />

Research, 30, 237e244.<br />

Campagna, C., Rivarola, M. M., Greene, D., & Tagliorette, A. (1995). Watching southern<br />

right whales in Patagonia. Unpublished report to UNEP, Nairobi.<br />

Constantine, R. (2001). Increased avoidance of swimmers by wild bottlenose<br />

dolphins (Tursiops truncatus) due to long-term exposure to swim-with-dolphin<br />

tourism. Marine Mammal Science, <strong>17</strong>(4), 689e702.<br />

Corkeron, P. J. (1995). Humpback whales (Megaptera novaeangliae) in Hervey Bay,<br />

Queensland: behaviour and responses to whale-watching vessels. Canadian<br />

Journal of Zoology, 73, 1290e1299.<br />

J.E.S. Higham, E.J. Shelton / Tourism Management 32 (2011) 1290e1298 1297<br />

Davis, H. (2002). Prediction and preparation: Pavlovian implications of research<br />

animals discriminating among humans. Institute for Laboratory Animal Research,<br />

43(1), 19e26.<br />

Davis, H., Wellwood, D., & Ciarniello, L. (2002). “Bear smart” community program:<br />

Background report. Victoria B.C.: British Columbia Ministry of Water, Land and<br />

Air Protection.<br />

Duffus, D. A., & Dearden, P. (1990). Non-consumptive wildlife-oriented recreation:<br />

a conceptual framework. Biological Conservation, 53(3), 213e231.<br />

Edds, P. L., & MacFarlane, J. A. F. (1987). Occurrence and general behaviour of<br />

balaenopterid cetaceans summering in the Saint Lawrence Estuary, Canada.<br />

Canadian Journal of Zoology, 65, 1363e1376.<br />

Edington, J. M., & Edington, M. A. (1986). Ecology, recreation and tourism. Cambridge:<br />

Cambridge University Press.<br />

Ellenberg, U., & Mattern, T. (2004). The most timorous of all? Impact of human<br />

disturbance on Humboldt penguins. New Zealand Journal of Zoology, 31,<br />

120e121.<br />

Ellenberg, U., Mattern, T., Seddon, P. J., & Jorquera, G. L. (2006). Physiological and<br />

reproductive consequences of human disturbance in Humboldt penguins: the<br />

need for species-specific visitor management. Biological Conservation, 133(1),<br />

95e106.<br />

Fennell, D. (2003). Ecotourism: An introduction (2nd ed.). London: Routledge.<br />

Gordon, J., Leaper, R., Hartley, F. G., & Chappell, O. (1992). Effects of whale watching<br />

vessels on the surface and underwater acoustic behaviour of sperm whales off<br />

Kaikoura, New Zealand. InScience and Research Series, Vol. 52. Wellington, New<br />

Zealand: Department of Conservation.<br />

Heatley, J. (1995). Antipredator conditioning in Mississippi Sandhill cranes (Grus<br />

canadensis pulla). Unpublished Master of Science, Texas A&M.<br />

Higginbottom, K. (Ed.). (2004). Wildlife tourism: Impacts, management and planning.<br />

Altona (Victoria, Australia): CRC for Sustainable Tourism.<br />

Higham, J. E. S. (1998). Tourists and albatrosses: the dynamics of tourism at the<br />

Northern Royal Albatross Colony, Taiaroa Head, New Zealand. Tourism<br />

Management, 19, 521e531.<br />

Higham, J. E. S., Bejder, L., & Lusseau, D. (2009). An integrated and adaptive<br />

management model to address the long-term sustainability of tourist interactions<br />

with cetaceans. Environmental Conservation, 35(4), 294e302.<br />

Higham, J. E. S., & Lusseau, D. (2004). Ecological impacts and management of tourist<br />

engagements with cetaceans. In R. Buckley (Ed.), Environmental impacts of<br />

ecotourism (pp. <strong>17</strong>3e188). Wallingford: CAB <strong>International</strong>.<br />

Higham, J. E. S., Lusseau, D., & Hendry, W. (2008). The viewing platforms from which<br />

animals are observed in the wild: a discussion of emerging research directions.<br />

Journal of Ecotourism, 7(2/3), 132e141, Special Issue on Australian Wildlife<br />

Tourism.<br />

Houston, D. M., & Russell, J. J. (2000). The impact of tourism on blue penguins<br />

(Eudyptula minor). Unpublished conference paper. The 4th international<br />

penguin conference, La Serena, Chile. September 4e8th 2000.<br />

Huber, L., Rechberger, S., & Taborsky, M. (2001). Social learning affects object<br />

exploration and manipulation in keas, Nestor notabilis. Animal Behaviour, 62,<br />

945e954.<br />

IWC. (2006). <strong>International</strong> whaling commission’s 58th annual meeting in St. Kitts and<br />

Nevis 2006. Retrieved 25 August, 2007, from. http://iwcoffice.org/_documents/<br />

commission/IWC58docs/iwc58docs.htm.<br />

Janik, V. M., & Thompson, P. M. (1996). Changes in surfacing patterns of<br />

bottlenose dolphins in response to boat traffic. Marine Mammal Science, 12,<br />

597e602.<br />

Johns, B. (1996). Responses of chimpanzees to habituation and tourism in the Kibale<br />

Forest, Uganda. Biological Conservation, 78, 257e262.<br />

Knight, J. (2009). Making wildlife viewable: habituation and attraction. Society and<br />

Animals, <strong>17</strong>, 167e184.<br />

Lusseau, D. (2003). The effects of tour boats on the behavior of bottlenose dolphins:<br />

using Markov chains to model anthropogenic impacts. Conservation Biology,<br />

<strong>17</strong>(6), <strong>17</strong>85e<strong>17</strong>93.<br />

Lusseau, D., & Higham, J. E. S. (2004). Managing the impacts of dolphin-based<br />

tourism through the definition of critical habitats: the case of bottlenose<br />

dolphins (Tursiops spp.) in Doubtful Sound, New Zealand. Tourism Management,<br />

25(5), 657e667.<br />

MacGibbon, J. (1991). Responses of sperm whales (Physeter macrocephalus) to<br />

commercial whale watching boats off the coast of Kaikoura. Unpublished report to<br />

the Department of Conservation. University of Canterbury, Christchurch, New<br />

Zealand.<br />

McKercher, B. (1993a). The unrecognised threat to tourism: can tourism survive<br />

“sustainability”? Tourism Management, 14(2), 131e136.<br />

McKercher, B. (1993b). Some fundamental truths about tourism: understanding<br />

tourism’s social and environmental impacts. Journal of Sustainable Tourism, 1(1),<br />

6e16.<br />

McKercher, B. (1998). The business of nature-based tourism. Melbourne: Hospitality<br />

Press.<br />

Mazur, J. (2006). Learing and behaviour (6th ed.). Upper Saddle River, NJ: Pearson<br />

Prentice Hall.<br />

Muloin, S. (1998). Wildlife tourism: the psychological benefits of whale watching.<br />

Pacific Tourism Review, 2, 199e212.<br />

National Park Service. (2006a). Deer and people at Fire Island National Seashore.<br />

Retrieved 27 March 2008, from. http://www.nps.gov.fiis/deerpeople/deer.html.<br />

National Park Service. (2006b). Deer and people at Fire Island National Seashore.<br />

Retrieved 22nd June, 2006, from. http://www.nps.gov/fiis/deerpeople/deer.<br />

html.


1298<br />

Newsome, D., Dowling, R., & Moore, S. (2005). Wildlife tourism. Clevedon: Channel View.<br />

Nimon, A. J., Schroter, R. C., & Stonehouse, B. (1995). Heart rate of disturbed<br />

penguins. Nature, 374(6521), 415.<br />

Nisbet, I. C. T. (2000). Disturbance, habituation, and management of waterbird<br />

colonies. Waterbirds, 23(2), 312e332.<br />

Nowacek, S. M., Wells, R. S., & Solow, A. R. (2001). Short-term effects of boat traffic<br />

on bottlenose dolphins, Tursiops truncatus, in Sarasota Bay, Florida. Marine<br />

Mammal Science, <strong>17</strong>(4), 673e688.<br />

Orams, M. (1995). Development and management of a feeding program for wild<br />

bottlenose dolphins at Tangalooma, Australia. Aquatic Mammals, 21(2), 137e147.<br />

Orams, M. (2000). Tourists getting close to whales, is it what whale-watching is all<br />

about? Tourism Management, 21(6), 561e569.<br />

van Polanen Petel, T. (2005). Effects of human activity on Weddell seals (Leptonychotes<br />

weddellii) in Antarctica. Retrieved 23 June, 2006, from. http://www.crctourism.<br />

com.au/CRCServer/documents/education/TamaraCompletion_report.doc.<br />

Romero, L., & Wikelsi, M. (2002). Exposure to tourism reduces stress-induced corticosterone<br />

levels in Galapagos marine iguanas. Biological Conservation, 108, 371e374.<br />

Salden, D. R. (1988). Humpback whales encounter rates offshore of Maui, Hawaii.<br />

Journal of Wildlife Management, 52(2), 301e304.<br />

Salvado, C. A. M., Kleiber, P., & Dizon, A. E. (1992). Optimal course by dolphins for<br />

detection avoidance. Fishery Bulletin, 90, 4<strong>17</strong>e420.<br />

J.E.S. Higham, E.J. Shelton / Tourism Management 32 (2011) 1290e1298<br />

Seddon, P., Ellenberg, U., & van Heezik, Y. (in press). Yellow-eyed penguin. In D.<br />

Boersma, & P. Garcia Borboroglu (Eds). Penguins. Washington, USA: University of<br />

Washington Press.<br />

Seddon, P., Smith, A., Dunlop, E., & Mathieu, R. (2004). Tourist visitor attitudes,<br />

activities and impacts at a yellow-eyed penguin breeding site on the Otago<br />

Peninsula, Dunedin, New Zealand. New Zealand Journal of Zoology, 31,<br />

119e120.<br />

Shackley, M. (1995). The future of gorilla tourism in Rwanda. Journal of Sustainable<br />

Tourism, 3(2), 61e72.<br />

Shackley, M. (1996). Wildlife tourism. London: Thomson Learning.<br />

Shelton, E. J., & Higham, J. E. S. (2007). Ecotourism and wildlife habituation. In<br />

J. E. S. Higham (Ed.), Critical issues in ecotourism: Understanding a complex<br />

tourism phenomenon (pp. 270e286). Oxford: Butterworth Heinemann.<br />

Shelton, E. J., Higham, J. E. S., & Seddon, P. (2004). Habituation, penguin research<br />

and ecotourism: some thoughts from left field. New Zealand Journal of Zoology,<br />

31(4), 119.<br />

Whittaker, D. (1997). Capacity norms on bear viewing platforms. Human Dimensions<br />

of Wildlife, 2(2), 37e49.<br />

Williams, R., Trites, A. W., & Bain, D. (2002). Behavioural responses of killer whales<br />

(Orcinus Orca) to whale-watching boats: opportunistic observations and<br />

experimental approaches. Journal of Zoology, 256, 255e270.


MARINE MAMMAL SCIENCE, 27(3): 606<strong>–</strong>621 (July 2011)<br />

C○ 2011 by the Society for Marine Mammalogy<br />

DOI: 10.1111/j.<strong>17</strong>48-7692.2010.00435.x<br />

Using distance sampling techniques to estimate<br />

bottlenose dolphin (Tursiops truncatus) abundance<br />

at Turneffe Atoll, Belize<br />

DOROTHY M. DICK 1<br />

ELLEN M. HINES<br />

Department of Geography and Human Environmental Studies,<br />

San Francisco State University,<br />

1600 Holloway Avenue,<br />

HSS Room 279,<br />

San Francisco, California 94132, U.S.A.<br />

E-mail: doridick14@gmail.com<br />

ABSTRACT<br />

Reliable abundance estimates are critical for management and conservation of<br />

coastal small cetaceans. This is particularly important in developing countries<br />

where coastal human populations are increasing, the impacts of anthropogenic<br />

activities are often unknown, and the resources necessary to assess coastal cetaceans<br />

are limited. We adapted ship-based line transect methods to small-boat surveys<br />

to estimate the abundance of bottlenose dolphins (Tursiops truncatus) at Turneffe<br />

Atoll, Belize. Using a systematic survey design with random start and uniform<br />

coverage, 34 dolphin clusters were sighted during small-boat line transect surveys<br />

conducted in 2005<strong>–</strong>2006. Distance sampling methods estimated abundance at 216<br />

individuals (CV = 27.7%, 95% CI = 126<strong>–</strong>370). Due to species rarity in the<br />

Atoll, small sample size, and potential violations in line transect assumptions, the<br />

estimate should be considered preliminary. Nevertheless, it provides up-to-date<br />

information on the status of a regional population in an area under increasing threat<br />

of habitat loss and prey depletion via uncontrolled development and unsustainable<br />

fishing. This information will be useful as Belize develops a new conservation<br />

initiative to create a comprehensive and resilient marine protected area system.<br />

Our study illustrates the application of distance sampling methods to small-boat<br />

surveys to obtain abundance estimates of coastal cetaceans in a region lacking<br />

resources.<br />

Key words: survey design, small-boat surveys, distance sampling, line transects,<br />

Distance 5.0, abundance, bottlenose dolphin, Tursiops truncatus, Turneffe Atoll,<br />

Belize.<br />

1 Also affiliated with: Oceanic Society, Fort Mason Quarters 35, San Francisco, California 94123,<br />

U.S.A. Current address: Department of Geosciences, Geography Program, Oregon State University, 104<br />

Wilkinson Hall, Corvallis, Oregon 97331, U.S.A.<br />

606


DICK AND HINES: BOTTLENOSE DOLPHIN ABUNDANCE 607<br />

Although the common bottlenose dolphin (Tursiops truncatus) is well studied and<br />

widespread across various marine habitats in temperate and tropical waters (Wells<br />

and Scott 1999), its worldwide status is unknown. In the IUCN 2008 Red List of<br />

Threatened Species update, T. truncatus was classified as Species of Least Concern and<br />

listed habitat destruction and degradation, disturbance and harassment, prey depletion,<br />

pollution, and direct/indirect takes as specific threats of concern (Hammond<br />

et al. 2008). While these threats apply to the species as a whole, any one may impact<br />

certain regional populations more than others as seen in the Mediterranean (Bearzi<br />

and Fortuna 2006, Bearzi et al. 2008), Moray Firth, Scotland (Wilson et al. 1997, Sini<br />

et al. 2005), and Peru (Van Waerebeek et al. 1997). Declines in regional populations<br />

of apex predators, such as bottlenose dolphins, could have far reaching effects on<br />

the community structure of an ecosystem (Currey et al. 2009), as seen in western<br />

Alaska when increased killer whale predation on sea otters lead to a dramatic rise<br />

in sea urchins and subsequent decreases in kelp forests (Estes et al. 1998). For these<br />

reasons, coastal bottlenose dolphin populations should be assessed on a regional scale<br />

when evaluating status, managing threats, and implementing conservation measures<br />

(Reeves et al. 2003, Currey et al. 2009).<br />

The effects of anthropogenic activities on coastal bottlenose dolphin populations<br />

are of particular concern as human populations continue to grow along coastlines,<br />

especially in developing countries (Aragones et al. 1997, Dawson et al. 2008). The<br />

extent of these activities, however, is generally unknown and the data needed to<br />

substantiate the impacts are rarely available. Without adequate knowledge about<br />

the status and life history of these populations future management actions are limited.<br />

Consequently, there is a strong need to conduct population assessment studies<br />

on coastal marine mammal populations in underdeveloped countries (Vidal 1993,<br />

Aragones et al. 1997, Hines et al. 2005, Dawson et al. 2008).<br />

Line transect surveys using distance sampling protocols are a common method<br />

used to assess marine mammal populations (Buckland et al. 2001). Surveys of this<br />

type typically use large ships or fixed-wing aircraft, sophisticated equipment (high<br />

powered binoculars, navigational tools, e.g., radar and electronic charts), and may traverse<br />

sizeable ocean expanses (Barlow 1988, 1995; Barlow et al. 1988; Calambokidis<br />

and Barlow 2004; Forcada et al. 2004; Mullin and Fulling 2004; Zerbini et al. 2007).<br />

As a result they are prohibitively expensive and inaccessible to developing countries<br />

with limited budgets and expertise and shallow coastal regions (Dawson et al. 2008).<br />

Adapting ship-based line transect methods to small-boat surveys (Vidal et al. 1997;<br />

Dawson et al. 2004; Williams and Thomas 2007, 2009) may be the best option to<br />

estimate coastal marine mammal populations in less affluent nations.<br />

Turneffe Atoll (Fig. 1), located in Belize, Central America, is the largest most<br />

biologically diverse of the nation’s three atolls (Stoddart 1962) and the only one<br />

without long-term ecological protection. Turneffe Atoll provides year-round habitat<br />

to a small population of coastal bottlenose dolphins (Grigg and Markowitz 1997,<br />

Campbell et al. 2002). Photo-identification studies from the 1990s estimated a<br />

population of less than 90 individuals (Campbell et al. 2002). More than a decade<br />

later there have been no new abundance estimates. Protected from import/export,<br />

wildlife trade, and hunting by Belize’s 1981 Wildlife Protection Act, threats from<br />

human induced mortality are currently minimal. However, unsustainable fishing<br />

(overfishing and illegal fishing) and rapid coastal development (mangrove clearing,<br />

dredging, and overdevelopment) have been identified as severe threats to the ecological<br />

integrity of Turneffe Atoll (World Resources Institute 2005, Granek 2006).<br />

There is a high probability that the small dolphin population could be threatened by


608 MARINE MAMMAL SCIENCE, VOL. 27, NO. 3, 2011<br />

Figure 1. The location of Turneffe Atoll, Belize, the survey design, and the bottlenose<br />

dolphin sighting locations during small-boat surveys conducted in 2005<strong>–</strong>2006. The survey<br />

was created with the automated survey design function in Program Distance 5.0 release 2<br />

using two substrata, the lagoon area and the western area between the mangrove cays and the<br />

fringing reef.<br />

habitat degradation, prey depletion, vessel traffic, and pollution as the atoll’s human<br />

population increases (Wells and Scott 1999). Obtaining an up-to-date abundance<br />

estimate for the population before further anthropogenic impacts occur is imperative<br />

to provide the baseline information necessary to guide future management and<br />

conservation actions.


DICK AND HINES: BOTTLENOSE DOLPHIN ABUNDANCE 609<br />

We sought to (1) develop and implement a repeatable and economically feasible<br />

systematic survey for use on bottlenose dolphins at Turneffe Atoll, (2) provide the<br />

first non-mark-recapture quantitative estimates of dolphin abundance in the study<br />

area, and (3) develop survey methods potentially applicable for other small cetacean<br />

coastal surveys in underdeveloped countries. This project is part of a long-term<br />

social and behavioral ecology research program of bottlenose dolphins conducted at<br />

Turneffe Atoll.<br />

MATERIALS AND METHODS<br />

Study Area<br />

Located 50 km east of mainland Belize and 18 km beyond the Mesoamerican Barrier<br />

Reef in the western Caribbean Sea, Turneffe Atoll is roughly 50 km long, 16 km<br />

wide at its widest point (average width 8<strong>–</strong>10 km), and encompasses approximately<br />

531 km2 (Fig. 1) (Stoddart 1962). The atoll contains >200 mangrove cays (islands),<br />

incorporates three shallow (


610 MARINE MAMMAL SCIENCE, VOL. 27, NO. 3, 2011<br />

Survey Procedures/Field Work<br />

Using the GPS unit to guide the boat along each transect line, surveys were<br />

conducted from either an 8.2 m skiff with two 85 hp outboard motors or a 6.4 m<br />

boat equipped with one 150 hp outboard motor. Boat speed was kept constant<br />

between 12 and 15 km/h and surveys were only conducted in Beaufort sea states of<br />

≤3 and wave heights


DICK AND HINES: BOTTLENOSE DOLPHIN ABUNDANCE 611<br />

one outlier at around 200 m, allowing better model fit (Buckland et al. 2001).<br />

Distance 5.0 release 2 (Thomas et al. 2010) was used to estimate detection probability<br />

using Conventional Distance Sampling (CDC) and Multiple Covariate Distance<br />

Sampling (MCDS) (Buckland et al. 2001, 2004). Unlike CDC, which assumes the<br />

detection of an object is the sole function of its distance from the line, MCDS allows<br />

for the inclusion of additional variables that are likely to impact the detection<br />

probability. For cetacean surveys, Beaufort sea state is a common covariate (Palka<br />

1996, Barlow et al. 2001) and was used here. Following Buckland et al. (2001),<br />

several standard detection function models in CDS and MCDS were fitted to the<br />

data. The Kolmogorov<strong>–</strong>Smirnov goodness-of-fit test for each model was used to examine<br />

the absolute fit of the model (Buckland et al. 2004). Estimates from the model<br />

with the lowest Akaike’s Information Criterion (AIC) were selected (Buckland et al.<br />

2001).<br />

Results from a size-biased regression (the default method in Distance) indicated<br />

weak evidence of dependence between cluster size and distance (rs = 0.19, P =<br />

0.14). Mean sample cluster size (¯s ) was considered an unbiased estimator of the<br />

estimated mean cluster size of the population ( Ê (s )) (or Ê (s ) = ¯s ) (Buckland et al.<br />

2001). Bottlenose dolphin abundance ( ˆ N) within the survey area was estimated using<br />

(Buckland et al. 2001, Marques and Buckland 2003):<br />

ˆN = A ·<br />

n · ¯s<br />

c · 2L · ,<br />

where ˆ N is the estimated abundance, A the size of the study area, n the number of<br />

clusters seen, ¯s the mean sample cluster size, c the constant (or multiplier) indicating<br />

the number of times each line was surveyed, L the total length of transect line, and<br />

the effective strip half-width.<br />

The coefficient of variations (CV) for n, , and ¯s were each calculated individually.<br />

Empirical estimation of CV(n), as recommended by Buckland et al. (2001), plus an<br />

adjustment for multiple visits to transect lines followed:<br />

where var(n) =<br />

CV(n) =<br />

TL<br />

var(n)<br />

n 2<br />

k<br />

ti · li · (n i/(ti · li) − n/TL) 2<br />

i=1<br />

k − 1<br />

TL is the total line length traveled, represented by, TL = k<br />

i=1 ti · li, ti the number<br />

of times a line was traveled, li the length of transect line i, ni the number<br />

of detections on line i, k the number of transect lines. CV() and CV(¯s ) were<br />

each calculated by dividing the standard error of the estimator by itself. Individual<br />

CVs were used to compute the abundance estimate CV using (Buckland et al.<br />

2001):<br />

CV( ˆ N) = [CV(n)] 2 + [CV()] 2 + [CV(¯s )] 2 .<br />

,


612 MARINE MAMMAL SCIENCE, VOL. 27, NO. 3, 2011<br />

RESULTS<br />

Survey Design<br />

The total planned survey effort was 220.54 km across 310.25 km2 . Due to shallow<br />

areas inaccessible to the boat, the realized survey effort was <strong>17</strong>5.21 km across<br />

298.5 km2 (about 80% of planned) for a set of transect lines. Six survey sets were<br />

completed, totaling 1,051.26 km of effort over 471.7 h during the rainy seasons of<br />

2005 (November<strong>–</strong>December) and 2006 (June<strong>–</strong>August and October<strong>–</strong>December) and<br />

during a two-week period (March<strong>–</strong>April) in the 2006 dry season.<br />

Abundance Estimation<br />

Thirty-four dolphin clusters, totaling 97 animals, were sighted on-effort (Fig. 1).<br />

Small sample size (n = 33 after truncation) precluded stratification by season; ungrouped<br />

perpendicular distances were pooled across all survey sets for analysis. Cluster<br />

size ranged from 1 to 12, with a mean cluster size of 2.61 (CV = <strong>17</strong>%, 95% CI =<br />

1.85<strong>–</strong>3.68). The majority of sightings (84.5%) were of clusters ≤3 dolphins.<br />

Model results and the best fitting plotted detection function are presented in<br />

Table 1 and Figure 2. Estimated abundance and density for the study area was<br />

216 dolphins (CV = 27.7%, 95% CI = 126<strong>–</strong>370) and 0.749 dolphins/km2 (CV =<br />

27.7%, 95% CI = 0.437<strong>–</strong>1.282), respectively. A detection probability of 0.50 (CV =<br />

15%, 95% CI = 0.372<strong>–</strong>0.681) occurred within the study area.<br />

DISCUSSION<br />

Our study produced a new abundance estimate for a species of conservation concern<br />

in a region where such estimates are scarce and without governmental financial<br />

support. Only two prior studies from the 1990s, both using photo-identification,<br />

Figure 2. Histogram of perpendicular sighting distances truncated at 90 m and the fitted<br />

detection function for the best fitting model, MCDS half-normal model (no adjustment<br />

parameters) with sea state as a covariate. One sighting at 200 m was truncated.


DICK AND HINES: BOTTLENOSE DOLPHIN ABUNDANCE 613<br />

Table 1. Model results for each of the tested models. Data were ungrouped, pooled across all surveys sets, and truncated at 90 m (n = 33).<br />

No. of parameters<br />

Model<br />

(Key + Adjustment) Key Adjustments AIC Ns<br />

ˆ CV(%) Dˆ CV(%) Ds<br />

ˆ CV(%) P CV(%) K-S p<br />

CDS<br />

Uniform + cosine 0 1 287.75 220 25.6 0.763 25.6 0.306 21.2 0.57 10.0 0.195<br />

Uniform + simple polynomial 0 1 288.04 204 25.0 0.707 25.0 0.272 20.5 0.64 9.0 0.153<br />

Half-normal + cosine 1 0 287.35 227 27.0 0.785 27.0 0.3<strong>17</strong> 22.8 0.55 13.0 0.210<br />

Half-normal + hermite 1 0 287.35 227 27.0 0.785 27.0 0.3<strong>17</strong> 22.8 0.55 13.0 0.210<br />

MCDS with Beaufort sea state as a covariate<br />

Half-normal + cosine 1 0 285.05 216 27.7 0.749 27.7 0.347 23.8 0.50 15.0 0.246<br />

Half-normal + hermite 1 0 285.05 216 27.7 0.749 27.7 0.347 23.8 0.50 15.0 0.246<br />

Hazard-rate + cosine 2 0 291.05 216 28.0 0.749 28.0 0.347 24.1 0.52 15.0 0.243<br />

Note: AIC = Akaike Information Criterion; ˆ Ns = abundance estimate for survey area; ˆ D = estimated density of dolphins/km 2 ; ˆ Ds = estimated<br />

density of clusters/km 2 ; P = detection probability; CV = coefficient of variation; K-S p = Kolmogorov<strong>–</strong>Smirnov goodness-of-fit p-value.


614 MARINE MAMMAL SCIENCE, VOL. 27, NO. 3, 2011<br />

report abundance estimates for regional bottlenose dolphin populations in Belize:<br />

Campbell et al. (2002) estimated 82<strong>–</strong>86 individuals at Turneffe Atoll and Kerr<br />

et al. (2005) reported 122 individuals in the Drowned Cayes, located 16 km west<br />

of Turneffe within the Mesoamerican Barrier Reef. The different dolphin abundance<br />

estimates for Turneffe is not surprising, primarily because the two methods measure<br />

slightly different things and are therefore, not directly comparable. Mark-recapture<br />

sampling estimates the abundance of the overall biological population whether or<br />

not all individuals are present at a moment in time, while line transect sampling<br />

estimates the size of the population within the study area during the survey interval<br />

(Calambokidis and Barlow 2004). Nevertheless, the abundance estimates for<br />

this region are small and suggest the population cannot withstand high levels of<br />

mortality.<br />

To our knowledge there are no reports of hunting, fisheries bycatch, or boat strikes<br />

of dolphins at Turneffe, and Campbell et al. (2002) reported an absence of crescentshaped<br />

scars indicative of shark predation. Moreover, bottlenose dolphins are fully<br />

protected under Belize’s 1981 Wildlife Protection Act. This would suggest threats to<br />

the dolphins at Turneffe are currently minimal. However, mounting evidence indicates<br />

that continued unsustainable fishing (overfishing and illegal fishing) and rapid<br />

development are impacting the ecological integrity of the Atoll (World Resources<br />

Institute 2005, Granek 2006). Within Belize, fisheries are primarily artisanal; ongoing<br />

commercial fisheries heavily and unsustainably exploit Caribbean spiny lobster<br />

and queen conch, and a small-scale fishery targets snapper, grouper, and other fish<br />

species (Gillet 2003, FAO 2005). Coblentz (1997) demonstrated artisanal fisheries<br />

to be unsustainable and can quickly alter reef communities. Furthermore, gill nets,<br />

recognized as a leading cause of cetacean mortality worldwide (Read et al. 2006),<br />

are one of several techniques used in this fishery, often by illegal Guatemalan and<br />

Honduran fishers (Gillet 2003, FAO 2005, Perez 2009). The threat of entanglement<br />

in gill nets is cause for concern to bottlenose dolphin populations in Turneffe and<br />

the rest of Belize.<br />

Removal of mangroves and dredging of seagrass beds for private and commercial<br />

development are increasing at Turneffe. As important nursery areas for reef fish<br />

(e.g., snappers, grunts, and parrotfish), elimination of either habitat type can have<br />

significant impacts to adjacent reef fish communities and may lead to cascading<br />

effects at higher trophic levels (Nagelkerken et al. 2002, Mumby et al. 2004, Manson<br />

et al. 2005). Although quantitative diet studies for Turneffe dolphins do not yet<br />

exist, if unsustainable fisheries and development are left unchecked, changes to<br />

the atoll’s ecosystem could potentially lead to prey depletion. This is especially<br />

concerning because females with neonatal and older, but presumed nursing, calves<br />

are sighted year-round suggesting the atoll may be an important calving area (Grigg<br />

and Markowitz 1997, Campbell et al. 2002). Prey depletion could have negative<br />

ramifications on reproductive success since lactating females have much higher energy<br />

requirements than non nursing individuals (Oftedal 1984, Cheal and Gales 1991).<br />

On a larger scale, site fidelity patterns indicate a high proportion of dolphins<br />

are transient and use a much larger area than Turneffe (Campbell et al. 2002).<br />

Belize is a small western Caribbean country and the dolphins observed at Turneffe<br />

could move beyond the country’s borders. Extensive gill net use occurs in Mexican,<br />

Honduran, and Guatemalan artisanal fisheries and bottlenose dolphins have been<br />

historically found entangled or dead in these nets (Vidal et al. 1994, FAO 2000).<br />

To better understand the impacts of these threats on Turneffe dolphins, surveys<br />

should continue at the atoll in conjunction with studies that focus on dolphin


DICK AND HINES: BOTTLENOSE DOLPHIN ABUNDANCE 615<br />

movement beyond the atoll, including a Belize-wide bottlenose dolphin population<br />

assessment, and the development of a method for monitoring and reporting fisheries<br />

interactions.<br />

Potential Violations of Line Transect Assumptions and Considerations for Future Surveys<br />

Survey design—Although it is possible to design a survey based on the amount<br />

of effort required to attain a certain level of precision (Buckland et al. 2001), often<br />

practical considerations will dictate the survey design, as was the case in this study.<br />

This is the first rigorous systematic survey conducted at Turneffe and its use is<br />

expected to continue. The choice of an equally spaced zig-zag design, with a few<br />

minor adjustments (see further), was considered successful.<br />

One particular issue arose during field surveys that required additional consideration.<br />

The full length of most transects between the mangrove shoreline and the<br />

fringing reef on the western side and several areas within the lagoons could not be<br />

surveyed; coral heads and shallow sand bars made these areas inaccessible. Instead,<br />

the areas were scanned for dolphins from the closest point that could be reached<br />

by the boat. Thomas et al. (2007) reported a similar experience and suggested this<br />

problem may be avoided by using high-resolution maps during the survey design<br />

process; however, such items are not available for this region. The issue was resolved<br />

by excluding the unsurveyed areas from the total line length traveled during analysis<br />

(Thomas et al. 2007). Because unsurveyed areas often occurred at the apexes of the<br />

zigzags, exclusion of these areas had the added benefit of further improving even<br />

coverage probability (Dawson et al. 2008).<br />

Abundance estimation—Visibility bias or incomplete detection at distance zero<br />

(g(0) < 1) can be problematic in marine mammal surveys and cause negatively<br />

biased estimates. This bias is classified in two ways: (1) animals may not be available<br />

to be seen by observers (availability bias) because they are not at the water’s surface<br />

where they can be seen (e.g., during diving), and (2) animals may potentially be<br />

visible to an observer but are not detected (perception bias) because of factors such as<br />

environmental conditions (e.g., sea state) (Marsh and Sinclair 1989). Consequently,<br />

species with longer dive times or found in smaller clusters are more likely to be<br />

missed by observers, thereby impacting the assumption g(0) = 1 (Dawson et al.<br />

2008). Bottlenose dolphins have relatively short dive durations in shallow areas. For<br />

example, along Florida’s Gulf Coast in areas with depths ranging between 1 and 5 m,<br />

average dive times for this species were recorded to be 20<strong>–</strong>25 s off Sanibel Island<br />

(Shane 1990), 30<strong>–</strong>40 s in Sarasota Bay (Irvine et al. 1981), and 28.5 s in Tampa Bay<br />

(Mate et al. 1995). Similar environmental conditions, including shallow waters, sand<br />

flats, and seagrass beds, occur at Turneffe Atoll; therefore dive times are likely to be<br />

comparable to those recorded in Florida. This short dive duration makes it unlikely<br />

that dolphins missed detection on the transect line because of being underwater. In<br />

addition, despite the small mean observed cluster size, almost 85% of the sightings<br />

were of clusters with ≤3 dolphins indicating smaller group sizes were not being<br />

missed. Introduction of significant bias to the abundance and density estimates from<br />

the assumption g(0) = 1 is, therefore, not expected.<br />

Responsive animal movement either toward or away from the survey vessel prior<br />

to detection will introduce bias (Turncock and Quinn 1991, Palka and Hammond<br />

2001). Bottlenose dolphins are known to bow-ride, and in at least 15% of the<br />

sightings, dolphins were first detected as they were quickly approaching the boat


616 MARINE MAMMAL SCIENCE, VOL. 27, NO. 3, 2011<br />

and just prior to bow-riding. Attraction to vessels is seen in other small cetacean<br />

species including Dall’s porpoise (Phocoenoides dalli, Turncock and Quinn 1991),<br />

Pacific white-sided dolphins (Lagenorhynchus obliquidens, Williams and Thomas 2007),<br />

and white-beaked dolphins (Lagenorhynchus albirostris, Palka and Hammond 2001).<br />

Lemon et al. (2006) noted behavioral changes by Indo-Pacific bottlenose dolphins<br />

(T. aduncus) to an approaching boat at 100 m. Calculated distances from this study<br />

found all but one sighting occurred at ≤90 m, suggesting that the presence of the<br />

boat may have already attracted dolphins to the boat prior to detection. During future<br />

surveys, every effort should be made to ensure dolphins are sighted as far ahead of<br />

the vessel as possible. Correction factors, to account for attractive movement and<br />

positively biased estimates, should be developed through the collection of animal<br />

orientation data (Palka and Hammond 2001).<br />

Measurement accuracy also influences estimates and can be problematic regardless<br />

of survey platform type, although it can be exacerbated in small-boat surveys. Estimation<br />

of distances and angles is common (Vidal et al. 1997, Hammond et al. 2002);<br />

however, it may lead to measurement rounding and biased estimates. For this reason,<br />

estimation is not recommended unless observers are well trained and continually<br />

tested throughout the survey or correction factors are developed (Buckland et al.<br />

2001, Dawson et al. 2008). Lerczak and Hobbs (1998) and Buckland et al. (2001)<br />

describe several acceptable methods for acquiring measurements using tools such as<br />

angle boards and binoculars with reticles and/or compasses. An angle board was used<br />

to measure sighting angles and it helped to avoid angle rounding. A nonstandard<br />

method was used to determine sighting distances due to site-specific limitations,<br />

primarily that the horizon was rarely visible due to the enclosed nature of the atoll.<br />

A clinometer was chosen, as it is self-leveling and does not require the horizon as a<br />

reference point. However, declination angles


DICK AND HINES: BOTTLENOSE DOLPHIN ABUNDANCE 6<strong>17</strong><br />

other small cetacean abundance studies (e.g., 27.9% short-beaked common dolphin<br />

(Delphinus delphis), 48.1% bottlenose dolphin (Barlow 1995); 15.7% Hector’s dolphin<br />

(Cephalorhynchus hectori) (Dawson et al. 2004); 29.2% Dall’s porpoise, 35.3% Pacific<br />

white-sided dolphin (Williams and Thomas 2007), the abundance (216 dolphins,<br />

CV = 27.7%) and density (0.749 dolphins/km 2 ,CV= 27.7%) reported here for<br />

Turneffe Atoll are the best estimates given the conditions and should be considered<br />

preliminary. The development of correction factors to help account for potential<br />

line transect analysis violations is strongly recommended. Future surveys would also<br />

benefit from a higher observation platform and the consistent use of binoculars;<br />

however, this should only be done after careful thought as protocol modification after<br />

six survey sets will impact the ability to detect trends.<br />

Being able to obtain reliable abundance estimates and identify critical habitats<br />

are vital to the creation and success of marine protected areas for cetaceans (Hoyt<br />

2005). Our results provide an up-to-date population assessment for the area and<br />

should serve as a baseline as Belize moves forward with the development of a new<br />

conservation initiative to create a comprehensive and resilient marine protected area<br />

system. The year-round presence of dolphins, including mom/calf pairs (Grigg and<br />

Markowitz 1997, Campbell et al. 2002), suggests this area may be an important<br />

calving/nursery site and should be a priority for protection. Moreover, the high<br />

frequency of sightings near mangrove shorelines and atoll openings, 2 suggests these<br />

areas are important habitat features to the dolphins and should remain undeveloped.<br />

This and other recent studies (Vidal et al. 1997; Dawson et al. 2004, 2008; Williams<br />

and Thomas 2009) have shown that line transect surveys can be successfully modified<br />

for small-boats, provided the survey design is well planned and field methods are<br />

designed to address the main assumptions of line transect analysis. As coastal human<br />

populations continue to grow and the threats to small cetacean species increase, the<br />

reduction in cost and the improved accessibility to shallow areas through the use<br />

of small-boats will provide more opportunities for underdeveloped nations to assess<br />

their coastal marine mammal populations.<br />

ACKNOWLEDGMENTS<br />

We thank L. Thomas for providing helpful guidance during the survey design phase,<br />

S. G. Allen, B. Holzman, and H. Edwards for their reviews of earlier drafts of this paper,<br />

M. Christman, C. Keller, and P. Kubilis for their statistical advice, and R. Williams and an<br />

anonymous reviewer for comments on the manuscript. We are grateful to the many Oceanic<br />

Society and Elderhostel volunteers who helped in the field. Financial and logistical support<br />

was provided by Oceanic Society Expeditions. D. Dick received financial support from San<br />

Francisco State University’s Department of Geography and the College of Behavioral and<br />

Social Sciences Graduate Stipend. All research was approved by the Office for the Protection<br />

of Human and Animal Subjects at SFSU and conducted under a research permit issued to the<br />

Oceanic Society from the Ministry of Natural Resources Forest Department, Belize.<br />

LITERATURE CITED<br />

Aragones, L. V., T. A. Jefferson and H. Marsh. 1997. Marine mammal survey techniques<br />

applicable in developing countries. Asian Marine Biology 14:15<strong>–</strong>39.<br />

2 1 Unpublished data provided by Dorothy Dick, 5227 /2 27th Avenue S, Gulfport, Florida 33707,<br />

July 2009.


618 MARINE MAMMAL SCIENCE, VOL. 27, NO. 3, 2011<br />

Barlow, J. 1988. Harbor porpoise, Phocoena phocoena, abundance estimation for California,<br />

Oregon, and Washington: I. ship surveys. Fishery Bulletin 86:4<strong>17</strong><strong>–</strong>432.<br />

Barlow, J. 1995. The abundance of cetaceans in California waters. Part 1: Ship surveys in<br />

summer and fall of 1991. Fishery Bulletin 93:1<strong>–</strong>14.<br />

Barlow, J., C. W. Oliver, T. D. Jackson and B. L. Taylor. 1988. Harbor porpoise, Phocoena phocoena,<br />

abundance estimation for California, Oregon, and Washington: II. aerial surveys.<br />

Fishery Bulletin 86:433<strong>–</strong>444.<br />

Barlow, J., T. Gerrodette and J. Forcada. 2001. Factors affecting perpendicular sighting<br />

distances on shipboard line-transect surveys for cetaceans. Journal of Cetacean Research<br />

and Management 3:201<strong>–</strong>212.<br />

Bearzi, G., and C. M. Fortuna. 2006. Common bottlenose dolphin Tursiops truncatus (Mediterranean<br />

subpopulation). Pages 64<strong>–</strong>73 in R. R. Reeves and G. N. di Sciara, eds. The status<br />

and distribution of cetaceans in the Black Sea and Mediterranean Sea. IUCN Centre for<br />

Mediterranean Cooperation, Malaga, Spain.<br />

Bearzi, G., S. Agazzi, S. Bonizzoni, M. Costa and A. Azzellino. 2008. Dolphins in a bottle:<br />

Abundance, residency patterns and conservation of bottlenose dolphins Tursiops truncatus<br />

in the semi-closed eutrophic Amvrakikos Gulf, Greece. Aquatic Conservation: Marine<br />

and Freshwater Ecosystems 18:130<strong>–</strong>146.<br />

Buckland, S. T., D. R. Anderson, K. P. Burnham, J. L. Laake, D. L. Borchers and L. Thomas.<br />

2001. Introduction to distance sampling: Estimating abundance of biological populations.<br />

Oxford University Press, Oxford, U.K.<br />

Buckland, S. T., D. R. Anderson, K. P. Burnham, J. L. Laake, D. L. Borchers and L. Thomas.<br />

2004. Advanced distance sampling: Estimating abundance of biological populations.<br />

Oxford University Press, Oxford, U.K.<br />

Calambokidis, J., and J. Barlow. 2004. Abundance of blue and humpback whales in the<br />

eastern North Pacific estimated by capture-recapture and line-transect methods. Marine<br />

Mammal Science 20:63<strong>–</strong>85.<br />

Campbell, G. S., B. A. Bilgre and R. H. Defran. 2002. Bottlenose dolphins (Tursiops truncatus)<br />

in Turneffe Atoll, Belize: Occurrence, site fidelity, group size, and abundance. Aquatic<br />

Mammals 28:<strong>17</strong>0<strong>–</strong>180.<br />

Cheal, A. J., and N. J. Gales. 1991. Body mass and food intake in captive, breeding bottlenose<br />

dolphins, Tursiops truncatus. Zoo Biology 10:451<strong>–</strong>456.<br />

Coblentz, B. E. 1997. Consumption of coral reef fish suggests non-sustainable extraction.<br />

Conservation Biology 11:559<strong>–</strong>561.<br />

Currey, R. J. C., S. M. Dawson and E. Slooten. 2009. An approach for regional threat<br />

assessment under IUCN Red List criteria that is robust to uncertainty: The Fiordland<br />

bottlenose dolphins are critically endangered. Biological Conservation 142:1570<strong>–</strong>1579.<br />

Dawson, S., E. Slooten, S. Dufresne and D. Clement. 2004. Small-boat surveys for coastal<br />

dolphins: Line-transect surveys for Hector’s dolphins (Cephalorhynchus hectori). Fisheries<br />

Bulletin 201:441<strong>–</strong>451.<br />

Dawson, S., P. Wade, E. Slooten and J. Barlow. 2008. Design and field methods for sighting<br />

surveys of cetaceans in coastal and riverine habitats. Mammal Review 38:19<strong>–</strong>49.<br />

ESRI. 2006. ArcInfo 9.2, 1999<strong>–</strong>2006. Environmental Systems Research Institute, Redlands,<br />

CA.<br />

Estes, J. A., M. T. Tinker, T. M. Williams and D. F. Doak. 1998. Killer whale predation on<br />

sea otters linking oceanic and nearshore ecosystems. Science 282:473<strong>–</strong>476.<br />

FAO. 2000. FAO fishery country profile: The Republic of Guatemala. Available at http://<br />

www.fao.org.fi/oldsite/FCP/en/GTM/profile.htm (accessed 14 March 2010).<br />

FAO. 2005. FAO fishery country profile: Belize. Available at http://www.fao.org/fi/<br />

oldsite/FCP/en/BLZ/profile.htm (accessed 14 March 2010).<br />

Forcada, J., M. Gazo, A. Aguilar, J. Gonzalvo and M. Fernández-Contreras. 2004. Bottlenose<br />

dolphin abundance in the NW Mediterranean: Addressing heterogeneity in distribution.<br />

Marine Ecology Progress Series 275:275<strong>–</strong>287.<br />

Gillet, V. 2003. The fisheries of Belize. Fisheries Centre Research Reports 11:141<strong>–</strong>147.


DICK AND HINES: BOTTLENOSE DOLPHIN ABUNDANCE 619<br />

Gischler, E. 2003. Holocene lagoonal development in the isolated carbonate platforms off<br />

Belize. Sedimentary Geology 159:113<strong>–</strong>132.<br />

Granek, E. F. 2006. Linkages between mangroves forests and coral reeds: quantifying disturbance<br />

effects and energy flow between systems. Ph.D. dissertation, Oregon State<br />

University, Corvalis, OR. 192 pp.<br />

Grigg, E., and H. Markowitz. 1997. Habitat use by bottlenose dolphins (Tursiops truncatus)<br />

at Turneffe Atoll, Belize. Aquatic Mammals 23:163<strong>–</strong><strong>17</strong>0.<br />

Hammond, P. S., P. Berggren, H. Benke, et al. 2002. Abundance of harbour porpoise and other<br />

cetaceans in the North Sea and adjacent waters. Journal of Applied Ecology 39:361<strong>–</strong>376.<br />

Hammond, P. S., G. Bearzi, A. Bjørge, et al. 2008. Tursiops truncatus. IUCN Red List of<br />

Threatened Species. Version 2009.2. Available at http://www.iucnredlist.org (accessed<br />

15 February 2010).<br />

Hines, E. M., K. Adulyanukosol and D. A. Duffus. 2005. Dugong (Dugong dugon) abundance<br />

along the Andaman Coast of Thailand. Marine Mammal Science 21:536<strong>–</strong>549.<br />

Hoyt, E. 2005. Marine protected areas for whales, dolphins and porpoises: A world handbook<br />

for cetacean habitat conservation. Earthscan, London, U.K.<br />

Irvine, A. B., M. D. Scott, R. S. Wells and J. H. Kaufmann. 1981. Movements and activities<br />

of the Atlantic bottlenose dolphin, Tursiops truncatus, near Sarasota, Florida. Fishery<br />

Bulletin 79:671<strong>–</strong>688.<br />

Kerr, K. A., R. H. Defran and G. S. Campbell. 2005. Bottlenose dolphins (Tursiops truncatus)in<br />

the Drowned Cayes, Belize: Group size, site fidelity, and abundance. Caribbean Journal<br />

of Science 41:<strong>17</strong>2<strong>–</strong><strong>17</strong>7.<br />

Lemon, M., T. P. Lynch, D. H. Cato and R. G. Harcourt. 2006. Response of traveling<br />

bottlenose dolphins (Tursiops aduncus) to experimental approaches by a powerboat in<br />

Jervis Bay, New South Wales, Australia. Biological Conservation. 127:363<strong>–</strong>372.<br />

Lerczak, J. A., and R. C. Hobbs. 1998. Calculating sightings distances from angular readings<br />

during shipboard, aerial, and shore-based marine mammal surveys. Marine Mammal<br />

Science 14:590<strong>–</strong>599.<br />

Manson, F. J., N. R. Loneragan, G. A. Skilleter and S. R. Phinn. 2005. An evaluation of<br />

the evidence for linkages between mangroves and fisheries: A synthesis of the literature<br />

and identification of research directions. Oceanography and Marine Biology: An Annual<br />

Review 43:485<strong>–</strong>515.<br />

Marques, F. F. C., and S. T. Buckland. 2003. Incorporating covariates into standard line<br />

transect analysis. Biometrics 59:924<strong>–</strong>935.<br />

Marsh, H., and D. F. Sinclair. 1989. Correcting for visibility bias in strip transect aerial<br />

surveys of aquatic fauna. Journal of Wildlife Management 53:10<strong>17</strong><strong>–</strong>1024.<br />

Mate, B. R., K. A. Rossbach, S. L. Nieukirk, R. S. Wells, A. B. Irvine, M. D. Scott and A. J.<br />

Read. 1995. Satellite-monitored movements and dive behavior of a bottlenose dolphin<br />

(Tursiops truncatus) in Tampa Bay, Florida. Marine Mammal Science 11:452<strong>–</strong>463.<br />

Minnesota Department of Natural Resources. 2008. DNR Garmin, version 5.1.1. The Minnesota<br />

Department of Natural Resources, St. Paul, MN. Available at http://www.<br />

dnr.state.mn.us/mis/gis/tools/arcview/extensions/DNRGarmin/DNRGarmin.html (accessed<br />

18 October 2008).<br />

Mullin, K. D., and G. L. Fulling. 2004. Abundance of cetaceans in the oceanic northern Gulf<br />

of Mexico, 1996<strong>–</strong>2001. Marine Mammal Science 20:787<strong>–</strong>807.<br />

Mumby, P. J., A. J. Edwards, J. E. Arias-González, et al.. 2004. Mangroves enhance the<br />

biomass of coral reef fish communities in the Caribbean. Nature 427:533<strong>–</strong>536.<br />

Nagelkerken, I., C. M. Roberts, G. Van Der Velde, M. Dorenbosch, M. C. van Riel, E.<br />

Cocheret de la Morinière and P. H. Nienhuis. 2002. How important for mangroves and<br />

seagrass beds for coral-reef fish? The nursery hypothesis tested on an island scale. Marine<br />

Ecology Progress Series 244:299<strong>–</strong>305.<br />

NASA Landsat Program. 2004. Landsat ETM+ Scene p018r048_7t20001108, Orthorectified,<br />

Geocover. EarthSat. Rockville, MD. Acquisition date 11/8/2000. Source for this data<br />

set Global land Cover Facility, University of Maryland, College Park, MD. Available at<br />

http://www.landcover.org (accessed 24 May 2005).


620 MARINE MAMMAL SCIENCE, VOL. 27, NO. 3, 2011<br />

Oftedal, O. T. 1984. Milk composition, milk yield and energy output at peak lactation:<br />

A comparative review. Pages 33<strong>–</strong>85 in M. Peaker, R. G. Vernon and C. H. Knight,<br />

eds. Physiological strategies in lactation. Symposia of the Zoological Society of London,<br />

number 51. Academic Press, London, U.K.<br />

Palka, D. 1996. Effects of Beaufort sea state on the sightability of harbor porpoises in the<br />

Gulf of Maine. Report of the <strong>International</strong> Whaling Commission 46:575<strong>–</strong>582.<br />

Palka, D. L., and P. S. Hammond. 2001. Accounting for responsive movement in line transect<br />

estimates of abundance. Canadian Journal of Fisheries and Aquatic Science 58:777<strong>–</strong>787.<br />

Perez, A. 2009. Fisheries management at the tri-national border between Belize, Guatemala,<br />

and Honduras. Marine Policy. 33:195<strong>–</strong>200.<br />

Platt, S. G., T. R. Rainwater, B. W. Miller and C. M. Miller. 2000. Notes on the mammals<br />

of Turneffe Atoll, Belize. Caribbean Journal of Science 36:166<strong>–</strong>168.<br />

Read, A. J., P. Drinker and S. Northridge. 2006. Bycatch of marine mammals in U.S. and<br />

global fisheries. Conservation Biology 20:163<strong>–</strong>169.<br />

Reeves, R. R., B. D. Smith, E. A. Crespo and G. N. di Sciara (compilers). 2003. Dolphins,<br />

whales and porpoises: 2002<strong>–</strong>2010 Conservation Action Plan for the World’s Cetaceans.<br />

IUCN/SSC Cetacean Specialist Group. IUCN, Gland, Switzerland.<br />

Shane, S. H. 1990. Behavior and ecology of the bottlenose dolphin at Sanibel Island, Florida.<br />

Pages 245<strong>–</strong>265 in S. Leatherwood and R. R. Reeves, eds. The bottlenose dolphin.<br />

Academic Press, San Diego, CA.<br />

Sini, M. I., S. J. Canning, K. A. Stockin and G. J. Pierce. 2005. Bottlenose dolphins around<br />

Aberdeen harbour, north-east Scotland: A short study of habitat utilization and the<br />

potential effects of boat traffic. Journal of the Marine Biological Association of the<br />

United Kingdom 85:1547<strong>–</strong>1554.<br />

Stoddart, D. R. 1962. Three Caribbean atolls: Turneffe Islands, Lighthouse Reef, and Glover’s<br />

Reef, British Honduras. Atoll Research Bulletin 87:1<strong>–</strong>151.<br />

Strindberg, S., and S. T. Buckland. 2004. Zigzag survey designs in line transect sampling.<br />

Journal of Agricultural, Biological, and Environmental Statistics 9:443<strong>–</strong>461.<br />

Thomas, L., S. T. Buckland, K. P. Burnham, D. R. Anderson, J. L. Laake, D. L. Borchers<br />

and S. Strindberg. 2002. Distance sampling. Encyclopedia of Environmetrics 1:544<strong>–</strong><br />

553.<br />

Thomas, L., R. Williams and D. Sandilands. 2007. Designing line transect surveys for complex<br />

survey regions. Journal of Cetacean Research and Management 9:1<strong>–</strong>11.<br />

Thomas, L., S. T. Buckland, E. A. Rexstad, J. L. Laake, S. Strindberg, S. Hedley, J. R. B.<br />

Bishop, T. A. Marques and K. P. Burnham. 2010. Distance software: Design and analysis<br />

of distance sampling surveys for estimating population size. Journal of Applied Ecology<br />

47:5<strong>–</strong>14.<br />

Turncock, B. J., and T. J. Quinn II. 1991. The effect of responsive movement on abundance<br />

estimation using line transect sampling. Biometrics 47:701<strong>–</strong>715.<br />

Van Waerebeek, K., M. F. Van Bressem, F. Felix, J. Alfaro-Shigueto, A. Garcia-Godos, L.<br />

Chávez-Lisambart, K. Ontón, D. Montes and R. Bello. 1997. Mortality of dolphins<br />

and porpoises in coastal fisheries off Peru and southern Ecuador in 1994. Biological<br />

Conservation 81:43<strong>–</strong>49.<br />

Vidal, O. 1993. Aquatic mammal conservation in Latin America: Problems and perspectives.<br />

Conservation Biology 7:788<strong>–</strong>795.<br />

Vidal, O., K. Van Waerebeek and L. T. Findley. 1994. Cetaceans and gillnet fisheries in<br />

Mexico, Central America and the wider Caribbean: A preliminary review. Report of the<br />

<strong>International</strong> Whaling Commission (Special Issue 15): 221<strong>–</strong>233.<br />

Vidal, O., J. Barlow, L. A. Hurtado, J. Torre, P. Cendón and Z. Ojeda. 1997. Distribution<br />

and abundance of the Amazon River dolphin (Inia geoffrensis) and the tucuxi (Sotalia<br />

fluviatilis) in the Upper Amazon River. Marine Mammal Science 13:427<strong>–</strong>445.<br />

Wells, R. S., and M. D. Scott. 1999. Bottlenose dolphin Tursiops truncatus (Montagu, 1821).<br />

Pages 137<strong>–</strong>182 in S. H. Ridgway and Sir R. Harrison, eds. Handbook of marine<br />

mammals. Volume 6. The second book of dolphins and the porpoises. Academic Press,<br />

San Diego, CA.


DICK AND HINES: BOTTLENOSE DOLPHIN ABUNDANCE 621<br />

Wilson, B., P. M. Thompson and P. S. Hammond. 1997. Habitat use by bottlenose dolphins:<br />

Seasonal distribution and stratified movement patterns in the Moray Firth, Scotland.<br />

Journal of Applied Ecology 34:1365<strong>–</strong>1374.<br />

Williams, R., and L. Thomas. 2007. Distribution and abundance of marine mammals in<br />

the coastal waters of British Columbia, Canada. Journal of Cetacean Research and<br />

Management 9:15<strong>–</strong>28.<br />

Williams, R., and L. Thomas. 2009. Cost-effective abundance estimation of rare animals: Testing<br />

performance of small-boat surveys for killer whales in British Columbia. Biological<br />

Conservation 142:1542<strong>–</strong>1547.<br />

Williams, R., R. Leaper, A. N. Zerbini and P. S. Hammond. 2007. Methods for investigating<br />

measurement error in cetacean line-transect surveys. Journal of the Marine Biological<br />

Association of the United Kingdom 87:313<strong>–</strong>320.<br />

World Resources Institute. 2005. Belize coastal threats atlas. 20 pp. Available at<br />

http://pdf.wri.org/belize_threat_atlas.pdf (accessed 14 March 2010).<br />

Zerbini, A. N., J. M. Waite, J. W. Durban, R. Leduc, M. E. Dahlheim and P. R. Wade. 2007.<br />

Estimating abundance of killer whales in the nearshore waters of the Gulf of Alaska and<br />

Aleutian Islands using line-transect sampling. Marine Biology 150:1033<strong>–</strong>1045.<br />

Received: 3 August 2009<br />

Accepted: 16 July 2010


Journal of Zoology<br />

The lesser of two evils: seasonal migrations of Amazonian<br />

manatees in the Western Amazon<br />

E. M. Arraut 1,2 , M. Marmontel 3 , J. E. Mantovani 4 , E. M.L.M. Novo 1 , D. W. Macdonald 2 & R. E. Kenward 5<br />

1 Earth Observation General Coordination, Remote Sensing Division, National Institute for Space Research, São José dos Campos, São Paulo, SP,<br />

Brazil<br />

2 Wildlife Conservation Research Unit, Zoology Department, The Recanati-Kaplan Centre, University of Oxford, Tubney, Oxfordshire, UK<br />

3 Mamirauá Sustainable Development Institute, Tefé, AM, Brazil<br />

4 National Institute for Space Research, Natal Regional Center, Natal, RN, Brazil<br />

5 Centre for Ecology and Hydrology, Wallingford, Oxfordshire, UK<br />

Keywords<br />

Trichechus inunguis; habitat selection;<br />

remote sensing; geographical information<br />

system; Sirenia; migration.<br />

Correspondence<br />

Eduardo Moraes Arraut, Earth Observation<br />

General Coordination, Remote Sensing<br />

Division, National Institute for Space<br />

Research, Avenida dos Astronautas, 1 758,<br />

Jardim da Granja, PO Box 515, CEP 12227-<br />

101, São José dos Campos, SP, Brazil. Tel:<br />

55 12 3945 6486; Fax: 55 12 3945 6488<br />

Email: arraut@dsr.inpe.br or<br />

arraut@gmail.com<br />

Editor: Andrew Kitchener<br />

Received 9 March 2009; revised 28 August<br />

2009; accepted 13 September 2009<br />

doi:10.1111/j.1469-7998.2009.00655.x<br />

Introduction<br />

Abstract<br />

Migrationisanadaptationtoenvironmentsinwhichhabitat<br />

quality in different regions changes asynchronously in space<br />

and/or time (Dingle & Drake, 2007). This implies that habitat<br />

quality in the destination will be better than that at the origin,<br />

but not necessarily that it must be good. Here, we show that<br />

Amazonian manatees that live in the region of the Mamirauá<br />

and Amana˜ Sustainable Development Reserves (RDSM and<br />

RDSA, respectively) are subject to challenging habitat conditions<br />

during part of the year, and that they migrate into an<br />

area that is their best option under difficult circumstances.<br />

Journal of Zoology. Print ISSN 0952-8369<br />

We investigated the paradox of why Amazonian manatees Trichechus inunguis<br />

undergo seasonal migrations to a habitat where they apparently fast. Ten males<br />

were tracked using VHF telemetry between 1994 and 2006 in the Mamirauá and<br />

Amana˜ Sustainable Development Reserves, constituting the only long-term<br />

dataset on Amazonian manatee movements in the wild. Their habitat was<br />

characterized by analysing aquatic space and macrophyte coverage dynamics<br />

associated with the annual flood-pulse cycle of the River Solimo˜es. Habitat<br />

information came from fieldwork, two hydrographs, a three-dimensional model<br />

of the water bodies and classifications of Landsat-TM/ETM + images. We show<br />

that during high-water season (mid-May to end-June), males stay in várzea lakes in<br />

association with macrophytes, which they select. We then show that, during lowwater<br />

(October<strong>–</strong>November), the drastic reduction in aquatic space in the várzea<br />

leads to the risk of their habitat drying out and increases the manatees’<br />

vulnerability to predators such as caimans, jaguars and humans. This explains<br />

why males migrate to Ria Amana˜. Based on data on illegal hunting, we argue that<br />

this habitat variability influences females to migrate too. We then use published<br />

knowledge of the environment’s dynamics to argue that when water levels are high,<br />

the habitats that can support the largest manatee populations are the várzeas of<br />

white-water rivers, and we conjecture that rias are the species’ main low-water<br />

refuges throughout Western Amazonia. Finally, we warn that the species may be<br />

at greater risk than previously thought, because migration and low-water levels<br />

make manatees particularly vulnerable to hunters. Moreover, because the flooding<br />

regime of Amazonian rivers is strongly related to large-scale climatic phenomena,<br />

there might be a perilous connection between climate change and the future<br />

prospects for the species. Our experience reveals that the success of research and<br />

conservation of wild Amazonian manatees depends on close working relationships<br />

with local inhabitants.<br />

We studied the influence of seasonal habitat variation on<br />

the migration of the Amazonian manatee Trichechus inunguis,<br />

which is the only member of the order Sirenia that lives<br />

exclusively in freshwater (Bertram & Bertram, 1973). Its<br />

distribution spans the Amazon basin, from Ecuador (Timm,<br />

Albuja & Clauson, 1986) and Peru (Reeves et al., 1996) to<br />

the Atlantic coast of Brazil (Best, 1984). Phylogenetic<br />

studies suggest that manatees from the mid-Solimo˜es, midand<br />

low-Amazonas form a single panmictic population<br />

(Cantanhede et al., 2005).<br />

Amazonian manatees are herbivores that in the RDSM<br />

and RDSA have been reported to feed on 63 species of<br />

Journal of Zoology 280 (2010) 247<strong>–</strong>256 c 2009 The Authors. Journal compilation c 2009 The Zoological Society of London 247


Seasonal migration of Amazonian manatees E. M. Arraut et al.<br />

aquatic macrophytes (annual freshwater plants) (Guterres &<br />

Marmontel, 2008). Adults reach up to 3 m in length and<br />

450 kg in weight (Caldwell & Caldwell, 1985), and consume<br />

about 8% of their body weight in aquatic macrophytes per<br />

day (Rosas, 1994). Because of strong habitat seasonality (see<br />

‘Materials and methods’), Amazonian manatees face an<br />

annual period of food shortage, which may last 7 months,<br />

during which they apparently fast (Best, 1983).<br />

The first tracking of an Amazonian manatee was carried<br />

out by Montgomery, Best & Yamakoshi (1981). A juvenile<br />

male was captured in the wild, kept in captivity for<br />

20 months (approximately half its life) and then released<br />

during the rising-water period in a várzea (a floodplain of a<br />

river with nutrient-rich and high-silt content water) near<br />

Manaus (a different region from where it had been captured).<br />

The manatee was tracked for 20 days, during which it<br />

spent most of its time feeding on aquatic macrophytes, and<br />

moved at a similar rate by day and by night.<br />

The Amazonian manatee’s migration was first reported<br />

by Marmontel et al. (2002), who tracked five adult males<br />

during a period of four and a half years. This showed that<br />

individuals migrated each year between várzea lakes, where<br />

they spent the high-water period, and Ria Amanã (a ria is<br />

long narrow lake formed by the partial submergence of a<br />

river valley), where they spent the low-water period (see<br />

‘Materials and methods’). Here, we extend the initial sample<br />

to 10 radio-tagged manatees to show (1) new migratory<br />

routes and (2) that the migration pattern remains consistent.<br />

We also indicate associations between Amazonian manatee<br />

248<br />

migratory movements and (3) availability of preferred food<br />

and (4) aquatic space reduction and predator aggregation.<br />

On that basis, we propose that the migratory behaviour,<br />

which is paradoxical insofar as animals travel to a place<br />

where, seemingly, they cannot eat, is a balance of feeding<br />

and predator avoidance. Finally, in the light of our results,<br />

we discuss the next steps for the species’ conservation.<br />

Materials and methods<br />

Study site<br />

The study region comprises just over 1 million hectares and<br />

lies within RDSM and RDSA, mid-River Solimo˜es region,<br />

Amazonas, Brazil (Fig. 1). The region was chosen because<br />

there were previous data on radio-tracked manatees (Marmontel<br />

et al., 2002), and the RDSM infrastructure facilitated<br />

fieldwork and the development of relationships with<br />

local communities. Its 120 km longitudinal extent<br />

(W65103 0 43.60 00 <strong>–</strong>W63151 0 49.80 00 ) encompassed all locations<br />

of the tracked manatees, and its 70 km latitudinal extent<br />

(S02114 0 39.34 00 <strong>–</strong>S03137 0 45.74) is bounded by várzea and by<br />

the Ria Amanã.<br />

In the várzea study region, water levels fluctuate annually<br />

over a range of 16 m, while in Ria Amana˜ annual variation is<br />

about 10 m (Fig. 2a). This variation in water level is caused<br />

by a flood pulse in the River Solimo˜es, and this pulse is the<br />

principal determinant of the extent and depth of flooding<br />

(Fig. 2b) and of the water’s physical chemistry (Junk, Bayley<br />

Figure 1 Study region showing in: small quadrant,<br />

location in relation to South America; large<br />

quadrant, band 4 of Landsat-TM image overlaid<br />

by migration routes of male Amazonian manatees<br />

Trichechus inunguis. Thin white lines are<br />

routes reported by Marmontel et al., (2002),<br />

thicker white lines are routes of manatees<br />

tracked later; thicker lines end at Lake Castanho<br />

so as not to mask previously detected<br />

routes, but all routes lead to or depart from Ria<br />

Amanã. Also shown are location of gauges<br />

(white squares) and of local human communities<br />

(white crosses).<br />

Journal of Zoology 280 (2010) 247<strong>–</strong>256 c 2009 The Authors. Journal compilation c 2009 The Zoological Society of London


E. M. Arraut et al.<br />

& Sparks, 1989). As a result of this large variation in water<br />

level, and because of the irregular geomorphology of many<br />

lakes and channels, water depth is the principal determinant<br />

of the connectivity between water bodies. The differences of<br />

geomorphology and of water level variation explain, for<br />

example, why although Lake Mamirauá and Ria Amana˜<br />

have similar depths during high-water, the former might<br />

become isolated during low-water while the latter remains<br />

deeper, connected to adjacent water bodies and with more<br />

aquatic space (Arraut, 2008). Based on this habitat seasonality,<br />

we distinguish between periods of lowering-water<br />

(July<strong>–</strong>August), low-water (September<strong>–</strong>November), risingwater<br />

(December<strong>–</strong>April) and high-water (May<strong>–</strong>June), defined<br />

on the basis of the water level and its speed of vertical<br />

change (Arraut, 2008).<br />

Manatee tracking<br />

Seasonal migration of Amazonian manatees<br />

Figure 2 (a) Water-level variation as a result of<br />

River Solimões’s flood pulse measured by<br />

gauges positioned in the channel of Lake<br />

Mamirauá and in Ria Amanã (see Fig. 1). The<br />

gauges are not inter-calibrated, so hydrographs<br />

are only comparable with respect to time and<br />

water-level variation. (b) Habitats vary depending<br />

on flooding duration and flood pulse phase.<br />

When water levels are high, lakes, rivers<br />

and flooded forest are all passable for the<br />

manatee. When levels are low, lakes may loose<br />

connectivity.<br />

From 1994 to 2006 (excepting 2004), 10 males were radiotracked<br />

with an Advanced Telemetry Systems (ATS) transmitter<br />

(Advanced Telemetry Systems Inc., Isanti, MN,<br />

USA) on the 164 MHz band using a three-element Yagi<br />

antenna. They were caught under government permit by<br />

research teams, sometimes aided by local hunters, mostly<br />

using a wide mesh net. Attempts were made to locate each<br />

individual once a day. Although hunting of Brazilian native<br />

fauna is illegal (Castello-Branco & Gomes, 1967), hunters<br />

eventually killed six out of our 10 tracked Amazonian<br />

manatees. We also captured three females, but we lost track<br />

of two on the following day and one was found dead 2 days<br />

later (Table 1). On some occasions, individuals were tracked<br />

Journal of Zoology 280 (2010) 247<strong>–</strong>256 c 2009 The Authors. Journal compilation c 2009 The Zoological Society of London 249


Seasonal migration of Amazonian manatees E. M. Arraut et al.<br />

Table 1 Information about tracked Amazonian manatees<br />

ID Gender<br />

Capture<br />

Location Water season<br />

but no GPS coordinate was associated to the location (Table<br />

1); in these instances, the geographical location is recorded<br />

as the name of the place (lake or part of lake) where the<br />

individuals were active and, for analysis, subsequently given<br />

coordinates at the centre of that location. As a result, for<br />

this subsample of the data, all locations of all manatees that<br />

had been recorded as being from that place were allocated<br />

the same coordinates. Thus, for example, individual seven<br />

was located 168 times, and these sites are allocated, for<br />

analysis, to only 12 GPS positions.<br />

Estimation of manatee locations<br />

We triangulated using the Best bi-angulation algorithm of<br />

the software LOCATION OF A SIGNAL (LOAS) 4.0b (ESS, 2006).<br />

Magnetic bearing corrections were calculated using the<br />

software DECLINAc¸a˜O MAGNÉTICA 2.0 (LEEE, 2003). Field<br />

experiments in which manatee tracking conditions were<br />

simulated at distances up to 2 km gave a mean positional<br />

error of 140 m (Arraut, 2008); we used a tracking resolution<br />

of 300 m, which was much smaller than either the range size<br />

or the dimensions of the habitat fragments.<br />

Migration detection and home-range<br />

calculations<br />

We first separated ranging from migratory movements using<br />

the dispersal detector algorithm in the software RANGES8<br />

(Kenward et al., 2008). We detected that all the eight<br />

manatees that we tracked for more than one flood-pulse<br />

period had migrated. Then, because in the scale of our<br />

analysis the habitat showed annual seasonality, (e.g. macro-<br />

Tracking period<br />

Seasonal<br />

ranges<br />

N_locs used for<br />

range estimate<br />

1 M L. Mamirauá High 4 years A 36 1289<br />

B 18 518<br />

2 M L. Mamirauá High 3 years 4 months A <strong>17</strong>0 794<br />

B 150 1952<br />

3 M L. Mamirauá High 4 years 4 months A 58 727<br />

B 50 1031<br />

4 M R. Amanã Lowering 1 year 4 month A 92 394<br />

B 61 2664<br />

5 M R. Amanã Low 5 months A 9 3474<br />

B 6 234<br />

6 M P. do Castanho Low 3 years A 85 308<br />

7 M L. Mamirauá High 2 years 2 months A 156 250<br />

8 M R. Amanã Lowering 3 months B 55 1186<br />

9 M R. Amanã Lowering 3 months B 43 21<strong>17</strong><br />

10 M R. Amanã Low 6 months B 19 748<br />

11 F P. do Castanho Lowering Not tracked <strong>–</strong> <strong>–</strong> <strong>–</strong><br />

12 F P. do Castanho Lowering Not tracked <strong>–</strong> <strong>–</strong> <strong>–</strong><br />

13 F R. Amanã Low Not tracked <strong>–</strong> <strong>–</strong> <strong>–</strong><br />

phyte cover and flooding were similar in high-water periods<br />

of different years), we pooled locations of each individual in<br />

the same geographic area and flood-pulse period. Home<br />

ranges were estimated using 95% kernels with fixed-smoothing<br />

parameter (Kenward et al., 2008). This model was<br />

especially appropriate because of variability in positional<br />

accuracy and owing to the small number of locations in<br />

some ranges (mean=59, range=6<strong>–</strong><strong>17</strong>0, n=15): kernels<br />

with fixed smoothing can give maximum range area estimates<br />

with as few as 12<strong>–</strong>15 locations (Kenward, 2001) (all<br />

but one animal had at least 12 locations); fixed smoothing<br />

also generates better overall surface estimates than adaptive<br />

smoothing (Seaman & Powell, 1996). Five individuals had<br />

two seasonal home ranges, whereas for the five other<br />

individuals, we could only calculate one home range.<br />

Habitat characterization<br />

95% kernel<br />

range size (ha)<br />

Number of locations used to calculate each of the seasonal home ranges is shown. Letter ‘A’ identifies high-water ranges in várzea lakes and<br />

letter ‘B’ identifies lowering-water ranges in River Japurá and low-water ranges in Ria Amanã (see ‘Habitat Analysis’ section).<br />

L., lake; P., paraná; M, male; F, female; ID, individual identification.<br />

250<br />

To analyse the spatial<strong>–</strong>temporal corollaries of manatee<br />

movement we considered: (1) water-level variation, measured<br />

daily from gauges with 1 cm interval positioned at<br />

Lake Mamirauá (since 1992) and Ria Amana˜ (since 2001);<br />

(2) the bottom topography and depths of water bodies,<br />

based on a bathymetric model from the field survey and<br />

geographical information system modelling; (3) the growth<br />

cycle and spatial distribution of macrophytes combined with<br />

the flooding dynamics in the area, from classifications of 11<br />

Landsat-TM and ETM+ images acquired between 1995<br />

and 2005 (three of the high-water, four of the loweringwater,<br />

two of the low-water and two of the rising-water).<br />

The total of 5 months of fieldwork was distributed throughout<br />

all four flood-pulse periods.<br />

Journal of Zoology 280 (2010) 247<strong>–</strong>256 c 2009 The Authors. Journal compilation c 2009 The Zoological Society of London


E. M. Arraut et al.<br />

Frequent cloud cover over the Amazon limited the availability<br />

of good-quality satellite images. Previous studies,<br />

however, suggest that on a broad scale, the environment<br />

shows annual seasonality with regard to the flooding extent,<br />

water physical chemistry and macrophyte cover dynamics<br />

(Junk et al., 1989; Junk & Piedade, 1993; Barbosa, 2005;<br />

Arraut, 2008). Thus, when images from the same dates as<br />

the tracking data were not available, we used images from a<br />

different year but same flood pulse period.<br />

In the classifications, we distinguished six habitat<br />

classes: (1) macrophytes on water; (2) flooded forest; (3)<br />

open-water; (4) macrophytes on non-flooded land; (5) nonflooded<br />

forest; (6) non-flooded land. The first three occurred<br />

during high-water and the last three, plus open-water,<br />

during the low-water. Classification assessment was<br />

based on a confusion matrix (Congalton & Green, 2008),<br />

which indicated accuracies above 90%. Further details of<br />

the habitat characterization process are given in Arraut<br />

(2008).<br />

Habitat analysis<br />

Habitat analysis addressed three questions:<br />

Question 1: Selective use of habitat<br />

Are males selective in their use of habitats, and specifically<br />

of aquatic macrophytes?<br />

Question 2: Proportion of forage within home ranges<br />

during high-water<br />

During the high-water, was there proportionally more<br />

forage in manatee home ranges within várzea lakes than if<br />

they had remained in River Japur a´ or Ria Amana˜?<br />

Question 3: Seasonal reduction in flooded area within<br />

home ranges<br />

Did the proportional seasonal reduction in the aquatic<br />

area within home ranges differ between manatees in várzea<br />

lakes and those in River Japurá or Ria Amana˜?<br />

Selective use of habitat (question 1) was assessed using<br />

compositional analysis (CA) on log ratios of used and<br />

available habitats (Aebischer, Robertson & Kenward,<br />

1993). To answer questions 2 and 3, we separated the 15<br />

home ranges occupied by the 10 manatees into two<br />

categories: ranges in várzea lakes during the high-water<br />

(category A) and ranges in River Japurá during the lowering-water<br />

or in Ria Amana˜ during the low-water<br />

(category B). Ranges in River Japur a´ and Ria Amana˜ were<br />

grouped into one category because the objective was to<br />

discover, with quantitative evidence, why male manatees<br />

remained in várzea lakes during high-water, and why they<br />

did not remain there during low-water (i.e. why they<br />

migrated); this grouping also increased sample size for<br />

statistical analyses.<br />

We carried out analyses of variance (ANOVA) by means<br />

of general linear models (GLM) using MINITAB v15.1.1<br />

(Minitab, 2007). For question 2, the response variable in<br />

the GLM was the ‘proportion of the home range covered by<br />

aquatic macrophytes’ (arcsine root transformed), and for<br />

question 3, it was the ‘proportional seasonal reduction in the<br />

flooded area’ (arcsine root transformed). Although there<br />

were too few ranges to test for inclusion of multiple<br />

predictor variables, we included individual identification<br />

(ID) and location number in each home range (N_locs) in<br />

the GLMs with category variable A/B to exclude the<br />

possibility that relationships with the response variable<br />

depended on differences between individuals and sampling<br />

characteristics.<br />

Results<br />

Seasonal habitat use and migration<br />

Males remained in várzea lakes from the middle of the rising<br />

to the beginning of the lowering-water period (approximately<br />

from February to the beginning of August), when<br />

flooding facilitated unconstrained access to all lakes, rivers<br />

and flooded forest. During the lowering-water, they left<br />

várzea lakes and migrated to Ria Amanã, where they stayed<br />

during the low-water. Then, during the rising-water, males<br />

migrated back to várzea lakes, thus closing the annual<br />

migratory cycle. The three females captured were either at<br />

(during low-water) or migrating to (during lowering-water)<br />

Ria Amana˜, thus, moving in accordance with what was<br />

observed for the radio-tracked males.<br />

Habitat analysis<br />

Selective use of habitat<br />

Seasonal migration of Amazonian manatees<br />

CA revealed aquatic macrophytes as the only preferred<br />

habitat class during high-water, and ranked the three<br />

habitat classes: (2) macrophytes, (1) open-water and (0)<br />

flooded forest [F (2, 5 d.f.)=7.56, P=0.04].<br />

Proportion of forage within home ranges during<br />

high-water<br />

During high-water, home ranges in várzea lakes (category<br />

A) encompassed almost sevenfold the area of aquatic<br />

macrophytes that occurred within home ranges in River<br />

Japurá or Ria Amana˜ (A: mean=0.27, SE=0.13; B:<br />

Figure 3 Macrophyte proportion in the home ranges of individuals<br />

was greater when in várzea lakes (^) than when in River JapuráorRia<br />

Amanã (’). ID, individual identification.<br />

Journal of Zoology 280 (2010) 247<strong>–</strong>256 c 2009 The Authors. Journal compilation c 2009 The Zoological Society of London 251


Seasonal migration of Amazonian manatees E. M. Arraut et al.<br />

mean=0.04, SE=0.04) (Fig. 3). In the GLM, the category<br />

A/B variable explained the most variance in the response<br />

variable, and continued to do so with the inclusion of ID<br />

and N_locs (Table 2). Model residuals showed no trend,<br />

indicating that the model was appropriate. Thus, the proportion<br />

of the home range classified as macrophyte habitat<br />

was greatest for males while they were in várzea lakes.<br />

Seasonal reduction in flooded area within home<br />

ranges<br />

Reduction in the flooded area in várzea lakes was over 4.5<br />

times greater than in River Japur a´ or Ria Amana˜ (A:<br />

mean=0.98, SE=0.01; B: mean=0.21, SE=0.07) (Fig. 4).<br />

In the GLM, variable A/B explained most of the variance in<br />

the response variable, and continued to do so with inclusion of<br />

ID and N_locs (Table 3). Analysis of full model residuals<br />

showed no particular trend, indicating the model was appropriate.<br />

Areas within high-water ranges practically dried out<br />

during low-water, while areas where manatees had low-water<br />

ranges suffered much smaller reduction in the aquatic space.<br />

After leaving Mamirauá while the water was lowering in<br />

1996, individuals one, two and three remained for a few days<br />

in an area of River Japur a. ´ They then disappeared during<br />

the low-water, only to return in the next rising-water. On the<br />

other hand, individuals captured at Paraná do Castanho<br />

during the lowering-water were moving in the direction of<br />

Table 2 General linear model results: F statistic and P values for<br />

response variable ‘macrophyte proportion in the home ranges’ for full<br />

and reduced models<br />

Explanatory variables A/B N_locs ID<br />

Variable alone 38.48 a<br />

0.51 d<br />

0.73 d<br />

A/B with each <strong>–</strong> 31.20 b<br />

67.50 b<br />

A/B with both <strong>–</strong> 121.36 c<br />

a Po0.001.<br />

b P=0.001.<br />

c P=0.002.<br />

d P not significant.<br />

ID, individual identification.<br />

Figure 4 Proportion reduction in flooded area (from high- to lowwater)<br />

was much greater for ranges in várzea lakes (^) than for those<br />

in River Japurá or Ria Amanã (’). ID, individual identification.<br />

252<br />

Table 3 General linear model results: F statistic and P values of the A/<br />

B variable ‘proportion reduction in flooded area’ for the full and<br />

reduced models<br />

Explanatory variables A/B N_locs ID<br />

Variable alone 227.37 a<br />

0.93 c<br />

0.40 c<br />

A/B with each <strong>–</strong> 238.94 a<br />

165.14 a<br />

A/B with both <strong>–</strong> 138.70 b<br />

a Po0.001.<br />

b P=0.001.<br />

c P not significant.<br />

ID, individual identification.<br />

Ria Amana˜, and all those captured and/or tracked during<br />

low-water were within Ria Amana˜ (n=9). This suggests<br />

that manatees aggregate in Ria Amana˜ during low-water.<br />

Discussion<br />

Várzea species are adapted to the seasonality of the environment<br />

(Junk et al., 1989). In the RDSM region, river dolphins<br />

Inia geoffrensis spend the high-water in várzea lakes and in<br />

flooded forest, but during low-water, they are in the main<br />

channels of Rivers Solimo˜es and Japur a ´ (Martin & da Silva,<br />

2004). Male Amazonian manatees also lived in várzea lakes<br />

during high-water, but in the low-water migrated to Ria<br />

Amana˜.<br />

The várzea lakes used during high-water offered forage in<br />

greater abundance [Results (1) and (2); Fig. 3, Table 2] and<br />

diversity (Junk et al., 1989; Arraut, 2008; Guterres &<br />

Marmontel, 2008), and less water current than rivers and<br />

paran as ´ (channels that connect rivers). When the water level<br />

dropped and the flooded area significantly contracted,<br />

restricting their home ranges [Result (3), Fig. 4, Table 3],<br />

male manatees migrated to Ria Amana˜ where they spent the<br />

low-water season. There, apart from more space, they<br />

encounter reduced risk from predators like: caimans Melanosuchus<br />

niger and Caiman crocodilus (Nunes Pereira, 1947),<br />

jaguars Panthera onca (Bertram & Bertram, 1973) and<br />

humans Homo sapiens sapiens (Domning, 1982a; Marmontel,<br />

2008). During the low-water period, caimans aggregate<br />

in whichever water bodies remain in the várzea (E. M.<br />

Arraut & M. Marmontel, pers. obs.), jaguars frequent the<br />

water margins (Ramalho, 2006) and humans come to gather<br />

the fish that aggregate there. Conversely, local hunters assert<br />

that the manatees in Ria Amana˜ are more difficult to kill<br />

(Calvimontes, 2009).<br />

Further information from interviews with local hunters<br />

indicates that female manatees have seasonal movement<br />

patterns similar to those of males. Sixty-four manatees were<br />

killed in Lake Castanho, Ria Amana˜ and the migration<br />

route between them during 2002<strong>–</strong>2004 (Calvimontes, 2009).<br />

These manatees were killed in Lake Castanho only between<br />

the rising- and the mid-lowering-water seasons and in Ria<br />

Amana˜, only between the lowering- and the beginning of the<br />

rising-water seasons. Of the 35 that had their sex determined,<br />

there were 13 males and 22 females. All four<br />

manatees killed in upper-Amanã at the beginning of the<br />

Journal of Zoology 280 (2010) 247<strong>–</strong>256 c 2009 The Authors. Journal compilation c 2009 The Zoological Society of London


E. M. Arraut et al.<br />

rising-water were females, including two that were calves<br />

(Calvimontes, 2009). From estimated birth dates of 24<br />

neonate Amazonian manatees, Best (1982) suggested that<br />

calves are normally born during the period of rising-water.<br />

As gestation is 12<strong>–</strong>14 months (Rosas, 1994), mating is<br />

expected to occur during low-water or the start of risingwater.<br />

Therefore, Ria Amana˜ may be both a mating and<br />

calving ground for Amazonian manatees.<br />

Do manatees that live in the várzeas of other<br />

white-water rivers move similarly as those in<br />

the mid-Solimões region?<br />

In the Amazon, immense white-water rivers such as the<br />

Solimo˜es, Amazonas, Puru´s and Madeira form várzeas and<br />

have rias annexed to their várzeas. They are also typified by<br />

flood pulses, and seasonal macrophytes (Junk & Furch,<br />

1993) (Fig. 5). The habitats available to manatees in these<br />

other regions of the Amazon, thus, seem similar to that<br />

which we have described in our study region.<br />

Evidence for the occurrence of manatees in these areas<br />

Acuna ˜ (1641), who reported them as abundant enough to be<br />

easily seen throughout the Rivers Solimo˜es and Amazonas,<br />

and Domning (1982a), who analysed the records of manatees<br />

hunted (legally at the time) that were documented in<br />

Manaus between <strong>17</strong>85 and 1983. Based on such information,<br />

as well as on how the large-scale movements of<br />

manatees in our study region coincide closely with the flood<br />

pulse, we conjecture that the habitats that can sustain the<br />

largest populations of T. inunguis during the high-water are<br />

the várzeas of large white-water rivers.<br />

On the other hand, during low-water, when most of the<br />

várzea dries out, rias remain the main places where aquatic<br />

space (Hess et al., 2003) with minimal current can be found.<br />

We thus conjecture that rias are the main low-water refuges<br />

of manatees that live in the floodplains of these major whitewater<br />

rivers; these floodplains comprise the great majority<br />

of the manatees’ known geographical range.<br />

Evolutionary ecology of the Amazonian<br />

manatees’ migration<br />

Seasonal migrations have been well documented in other<br />

aquatic mammals, and the ecological processes that drive<br />

River Japurá<br />

Ria Tefé<br />

Ria Amanã<br />

River Solimões<br />

River Negro<br />

River Amazonas<br />

migration often seem to be related to habitat variability. The<br />

Florida manatee Trichechus manatus latirostris (Deutsch<br />

et al., 2003) and the dugong Dugong dugon (Sheppard et al.,<br />

2006) migrate to warmer waters in autumn, and back to<br />

their summer areas in spring. In these two species, the<br />

evolution of migration can be explained as a means of<br />

optimizing forage when it is available and avoiding harsh<br />

environmental conditions during the rest of the year. We<br />

have demonstrated comparable seasonal movements for<br />

Amazonian manatees, although in their case the ‘seasons’<br />

are determined by variation in water level and the limiting<br />

factors are space and predation.<br />

Trichechus inunguis is thought to be the most derived<br />

species of manatee, having evolved from ancestral trichechids<br />

that colonized the Western Amazon in the Pliocene<br />

(2<strong>–</strong>6 million years ago) (Domning, 1982b). Geological evidence<br />

suggests that the várzea of River Solimo˜es achieved its<br />

present form in the Holocene (o10 000 years), and that<br />

terraces on its sides were formed a bit earlier, in the last<br />

47 300 years (Rossetti, Toledo & G oes, ´ 2005). Because rias<br />

are formed by the excavation of such terraces by ancient<br />

rivers, they must be younger than the terraces. Thus, the use<br />

of Ria Amana˜ as a breeding and calving ground, and<br />

migration to it, must be a relatively recent phenomenon in<br />

the species’ evolutionary history and, if our conjecture is<br />

correct, this also applies to the use of other rias. The<br />

geological history of the Western Amazon thus indicates<br />

that Amazonian manatees have changed their spatial behaviour<br />

in response to changes in the habitat that occurred<br />

within the last few tens of thousands of years.<br />

The fact that manatees living from the mouth of the River<br />

Amazon and those occurring throughout Amazonia to the<br />

west are part of a single panmictic population (Cantanhede<br />

et al., 2005) indicates that any differences in the movements<br />

of individuals in these widely separated areas are probably a<br />

result of behavioural plasticity in response to local habitat<br />

conditions. This raises the question: what are the movement<br />

patterns of the manatees living near the mouth of River<br />

Amazon (Domning, 1981), where water level varies daily<br />

(ANA, 2009) because of the Atlantic tides and where there<br />

are no rias? How do these differences in the spatial<strong>–</strong>temporal<br />

dynamics of their habitats affect manatee reproductive<br />

ecology? For example, if manatees find forage and living<br />

space in the same area throughout the year, we would not<br />

River Tapajós<br />

N<br />

1320 km<br />

Seasonal migration of Amazonian manatees<br />

Figure 5 Rias (shown by arrows), which are<br />

numerous and of various sizes, occur throughout<br />

the margins of Rivers Solimões and Amazonas<br />

and are conjectured to be the Amazonian<br />

manatee’s main low-water refuges. Study-region<br />

limits indicated by dashed-line polygon.<br />

Landsat-TM mosaic of the central Amazon<br />

várzea. Source: Shimabukuro, Novo & Mertes<br />

(2002).<br />

Journal of Zoology 280 (2010) 247<strong>–</strong>256 c 2009 The Authors. Journal compilation c 2009 The Zoological Society of London 253


Seasonal migration of Amazonian manatees E. M. Arraut et al.<br />

expect them to migrate. Answers to these questions await<br />

further research.<br />

Conservation implications<br />

Migrating Amazonian manatees pass through narrow channels<br />

where hunters can wait within harpoon range. Because<br />

human settlements are numerous and widespread (Fig. 1),<br />

and local inhabitants relish manatee meat, and are aware of<br />

the timing of their migration, manatees have to pass through<br />

perilous bottlenecks to arrive at the relative safety of Ria<br />

Amana˜. Then, in the beginning of the rising-water, water<br />

levels rise sufficiently to flood the few beaches where macrophytes<br />

have been growing close to the mouth of Ria Amana˜.<br />

The manatees aggregate at these beaches, hungry after<br />

several months of fasting and again become vulnerable to<br />

hunters. Hence, the migration to, and prolonged occupation<br />

of, Ria Amana˜ present clear risks to contemporary manatees.<br />

Nevertheless, the fact that manatees migrate to Ria<br />

Amana˜, and that, as argued above, this migratory behaviour<br />

is probably not genetically determined, suggests that staying<br />

in várzea lakes during the low-water has been (at least until<br />

recently) even more perilous. The lowering- and low-water<br />

seasons are, thus, periods when Amazonian manatees living<br />

in the mid-Solimo˜es region have to choose between the<br />

lesser of two evils.<br />

Any climate changes that lead to more intense droughts,<br />

and thus to a reduction in the aquatic space during the lowwater<br />

period, would increase the exposure of Amazonian<br />

manatees to hunters. A connection between climate change<br />

and an increase in the frequency of droughts in the Amazon<br />

is predicted under all scenarios of all 23 IPCC climatic<br />

models (Malhi et al., 2008), and relationships between<br />

large-scale climatic phenomena, such as the El Nino ˜ Southern<br />

Oscillation and the Tropical North Atlantic Ocean<br />

Temperatures, and the levels of major Amazonian rivers,<br />

were determined by Richey, Nobre & Deser (1989), Schongart<br />

& Junk (2007) and Marengo et al. (2008a,b). A situation<br />

that might become more frequent is exemplified by the<br />

drought of 2005, estimated to be the most intense in the last<br />

40 years, when the Amazon River and its major tributaries<br />

were left with only a fraction of their normal volume<br />

(Marengo et al., 2008a,b). During this drought, enormous<br />

quantities of fish died, clogging rivers and poisoning the<br />

water, and people living in small communities had to walk<br />

several kilometres to find water and food (Giles, 2006).<br />

Amazonian manatees also suffered. Local people informed<br />

us of the killings of at least five manatees in Ria Tefe´, and<br />

colleagues working in River Puru´s informed us of at least 12<br />

that were killed in Ria Jari (B. Marioni, E. V. Mullen & F.<br />

Rossoni, pers. comm.).<br />

In order to protect the manatees better now, and to<br />

safeguard them in the face of climate change, it is important<br />

to research their status in other regions, and to complement<br />

such research with education. Our experience indicates that to<br />

be effective, any conservation programme needs to learn from,<br />

educate and bring benefits to local populations, because local<br />

people are necessary actors in conserving the manatees and<br />

254<br />

their environment. Our conjecture that rias are the manatee’s<br />

main low-water refuge can serve as a starting point in the<br />

search for manatees in other regions of the Amazon.<br />

Acknowledgements<br />

We are grateful to the Coordenac¸ão de Aperfeiçoamento de<br />

Pessoal de Nível Superior (CAPES) for funding Arraut’s<br />

PhD stipend and expenses, to the Rede de Modelagem<br />

Ambiental da Amazoˆnia (GEOMA) Network and to the<br />

Graduate program in Remote Sensing at INPE (Brazil). We<br />

are thankful to Antonio Pinto de Oliveira for his help during<br />

fieldwork, and to all the local hunters and trainees who<br />

helped with data collection. We thank Dr Paul Johnson for<br />

his suggestions during data analyses, and Drs Marion<br />

Valeix, Jorgelina Marino, Thomas Merckx, Laura Fasola<br />

and Ruth Kanski for fruitful comments on an earlier draft.<br />

We also thank Dr Dilce Rossetti, Dr Thiago Silva and two<br />

anonymous referees for their suggestions.<br />

References<br />

Acuna, ˜ C. (1641). Novo descobrimento do Grande Rio Amazonas/Critóbal<br />

de Acuña, 1641. Rio de Janeiro: Agir.<br />

Aebischer, N.J., Robertson, P.A. & Kenward, R.E. (1993).<br />

Compositional analysis of habitat use from animal radiotracking<br />

data. Ecology 74, 1313<strong>–</strong>1325.<br />

ANA. (2009). Hidroweb <strong>–</strong> Hydrological Information System,<br />

Brazilian National Agency of the Waters. Available at<br />

http://hidroweb.ana.gov.br/ (accessed 10 February 2009).<br />

Arraut, E. (2008). Migraça˜o do Peixe-boi Amazoˆnico: uma<br />

abordagem por sensoriamento remoto, radiotelemetria e<br />

geoprocessamento. PhD thesis, Instituto Nacional de Pesquisas<br />

Espaciais (National Institute for Space Research),<br />

Sa˜o Jose´ dos Campos.<br />

Barbosa, C.C.F. (2005). Sensoriamento Remoto da dinaˆmica<br />

da circulaça˜o daágua do sistema planície de Curuai/Rio<br />

Amazonas. Sa˜oJose´ dos Campos: Instituto Nacional de<br />

Pesquisas Espaciais.<br />

Bertram, G.C.L. & Bertram, C.K.R. (1973). The modern<br />

Sirenia: their distribution and status. Biol. J. Linn. Soc. 5,<br />

297<strong>–</strong>338.<br />

Best, R.C. (1982). Seasonal breeding in the Amazonian<br />

manatee, Trichechus inunguis (Mammalia, Sirenia). Biotropica<br />

14, 76<strong>–</strong>78.<br />

Best, R.C. (1983). Apparent dry-season fasting in Amazonian<br />

Manatees (Mammalia: Sirenia). Biotropica 15, 61<strong>–</strong>64.<br />

Best, R.C. (1984). The aquatic mammals and reptiles of the<br />

Amazon. In The Amazon limnology and landscape ecology<br />

of a mighty tropical river and its basins: 371<strong>–</strong>412. Sioly, H.<br />

(Ed.). Dordrecht: Dr Junk W. Publishers.<br />

Caldwell, D.K. & Caldwell, M.C. (1985). Manatees Trichechus<br />

manatus (Linnaeus, <strong>17</strong>58); Trichechus senegalensis<br />

Journal of Zoology 280 (2010) 247<strong>–</strong>256 c 2009 The Authors. Journal compilation c 2009 The Zoological Society of London


E. M. Arraut et al.<br />

(Link, <strong>17</strong>95) and Trichechus inunguis (Natterer, 1883). In<br />

Handbook of marine mammals: 33<strong>–</strong>66. Ridgway, S.H. &<br />

Harisson, R. (Eds). London: Academic Press.<br />

Calvimontes, J. (2009). Etnoconocimiento, uso y conservación<br />

del manatí amazónico Trichechus inunguis en la Reserva de<br />

Desarrollo Sostenible Amana˜, Brasil. Lima: Universidad<br />

Nacional Agraria La Molina.<br />

Cantanhede, A., Da Silva, V., Farias, I., Hrbek, T., Lazzarini,<br />

S. & Alves-Gomes, J. (2005). Phylogeography and population<br />

genetics of the endangered Amazonian manatee,<br />

Trichechus inunguis Natterer, 1883 (Mammalia, Sirenia).<br />

Mol. Ecol. 14, 401<strong>–</strong>413.<br />

Castello-Branco, H. & Gomes, S.F. (1967). Lei de proteça˜oda<br />

fauna silvestre (Law for the protection of native fauna).<br />

Brasília: C. C. Presideˆncia da Repu´blica, Subchefia de<br />

Assuntos Jur ıdicos. ´<br />

Congalton, R.G. & Green, K. (2008). Assessing the<br />

accuracy of remotely sensed data: principles and practices,<br />

2nd edn. Boca Raton: CRC Press, Taylor & Francis<br />

Group.<br />

Deutsch, C., Reid, J., Bonde, R., Easton, D., Kochman, H. &<br />

O’Shea, T. (2003). Seasonal movements, migratory behavior,<br />

and site fidelity of West Indian manatees along the<br />

Atlantic coast of the United States. Wildl. Monogr. 151,<br />

1<strong>–</strong>77.<br />

Dingle, H. & Drake, V. (2007). What is migration? Bioscience<br />

57, 113<strong>–</strong>121.<br />

Domning, D. (1981). Distribution and status of manatees<br />

Trichechus spp. near the mouth of the Amazon River,<br />

Brazil. Biol. Conserv. 19, 85<strong>–</strong>97.<br />

Domning, D. (1982a). Commercial exploitation of manatees<br />

Trichechus in Brazil c. <strong>17</strong>85<strong>–</strong>1973. Biol. Conserv. 22,<br />

101<strong>–</strong>126.<br />

Domning, D. (1982b). Evolution of manatees: a speculative<br />

history. J. Paleontol. 56, 599<strong>–</strong>619.<br />

ESS. (2006). Location of a Signal (LOAS). Heygmagas:<br />

Ecological Software Solutions LLC.<br />

Giles, J. (2006). The outlook for Amazonia is dry. In Nature<br />

news: 726<strong>–</strong>727. Campbell, P. & Lincoln, T. (Eds). London:<br />

Nature Publishing Group.<br />

Guterres, M. & Marmontel, M. (2008). Anatomia e Morfologia<br />

de plantas aquáticas da Amazoˆnia utlizadas como potencial<br />

alimento pelo peixe-boi amazoˆnico. Bele´m: Instituto de<br />

Desenvolvimento Sustentável Mamirau a. ´<br />

Hess, L.L., Melack, J.M., Novo, E., Barbosa, C.C.F. &<br />

Gastil, M. (2003). Dual-season mapping of wetland inundation<br />

and vegetation for the central Amazon basin.<br />

Remote Sens. Environ. 87, 404<strong>–</strong>428.<br />

Junk, W.J., Bayley, P.B. & Sparks, R.E. (1989). The flood<br />

pulse concept in river<strong>–</strong>floodplain systems. In Proceedings of<br />

the <strong>International</strong> Large River Symposium (LARS):<br />

110<strong>–</strong>127. Dodge, D.P. (Ed.). Toronto: Canadian Special<br />

Publication of Fisheries and Aquatic Sciences.<br />

Junk, W.J. & Furch, K. (1993). A general review of tropical<br />

South American floodplains. Wetl. Ecol. Mgmt. 2,<br />

231<strong>–</strong>238.<br />

Seasonal migration of Amazonian manatees<br />

Junk, W.J. & Piedade, M.T.F. (1993). Biomass and<br />

primary-production of herbaceous plant communities<br />

in the Amazon floodplain. Hydrobiologia 263,<br />

155<strong>–</strong>162.<br />

Kenward, R.E. (2001). A manual for wildlife radio tagging.<br />

London: Academic Press.<br />

Kenward, R.E., Walls, S.S., South, A.B. & Casey, N. (2008).<br />

Ranges8 for the analysis of tracking and location data.<br />

Wareham: Anatrack Ltd.<br />

LEEE. (2003). Declinaça˜o Magne´tica 2.0. Santa Catarina:<br />

Laborat orio ´ de Eficieˆncia Energética em Edificac¸o˜es.<br />

Malhi, Y., Roberts, J.T., Betts, R.A., Killeen, T.J., Li, W. &<br />

Nobre, C.A. (2008). Climate change, deforestation, and the<br />

fate of the Amazon. Science 319, 169<strong>–</strong><strong>17</strong>2.<br />

Marengo, J.A., Nobre, C.A., Tomasella, J., Cardoso, M.F. &<br />

Oyama, M.D. (2008a). Hydro-climatic and ecological<br />

behaviour of the drought of Amazonia in 2005. Philos.<br />

Trans. Roy. Soc. Lond. Ser. B Biol. Sci. 363, <strong>17</strong>73<strong>–</strong><strong>17</strong>78.<br />

Marengo, J.A., Nobre, C.A., Tomasella, J., Oyama, M.D.,<br />

Oliveira, G.S., Oliveira, R., Alves, L.M. & Brown, I.F.<br />

(2008b). The drought of Amazonia in 2005. J. Clim. 21,<br />

495<strong>–</strong>516.<br />

Marmontel, M. (2008). Trichechus inunguis. Available at<br />

http://www.iucnredlist.org (accessed 19 February 2009).<br />

Marmontel, M., Guterres, M., Meirelles, A.C., Calvimontes,<br />

J. & Rosas, F.C.W. (2002). Lago Amana˜: destino estival de<br />

manaties amazoˆnicos en la Amazonia occidental<br />

brasilena.: ˜ 40. Reunion de Trabajo de Especialistas en<br />

Mam ıferos ´ Acuáticos de Ame´rica del Sur (RT), 10<br />

Congresso Sociedad Latinoamericana de Especialistas<br />

en Mam ıferos ´ Acuáticos (SOLAMAC), Vald ıvia, ´<br />

Chile.<br />

Martin, A.R. & da Silva, V.M.F. (2004). River dolphins and<br />

flooded forest: seasonal habitat use and sexual segregation<br />

of botos (Inia geoffrensis) in an extreme cetacean environment.<br />

J Zool. (Lond.) 263, 295<strong>–</strong>305.<br />

Minitab. (2007). Minitab. State College: Minitab Inc.<br />

Montgomery, G.G., Best, R.C. & Yamakoshi, M. (1981). A<br />

radio-tracking study of the Amazonian manatee<br />

Trichechus inuguis (Mammalia: Sirenia). Biotropica 13,<br />

81<strong>–</strong>85.<br />

Nunes Pereira, M. (1947). O peixe-boi da Amazoˆnia. Manaus:<br />

Imprensa Official.<br />

Ramalho, E.E. (2006). Uso do hábitat e dieta da Onça-pintada<br />

(Panthera onca) em uma área de várzea, Reserva de<br />

Desenvolvimento Sustentável Mamirauá, Amazoˆnia Central,<br />

Brasil. Masters thesis, Instituto Nacional de Pesquisas da<br />

Amazoˆnia (National Institute for Research in Amazonia),<br />

Manaus.<br />

Reeves, R.R., Leatherwood, S., Jefferson, T.A. & Curry, B.E.<br />

(1996). Amazonian manatees, Trichechus inunguis, in Peru:<br />

distribution, exploitation, and conservation status. Intercieˆncia<br />

21, 246<strong>–</strong>254.<br />

Richey, J.E., Nobre, C. & Deser, C. (1989). Amazon river<br />

discharge and climate variability: 1903 to 1985. Science<br />

246, 101<strong>–</strong>103.<br />

Journal of Zoology 280 (2010) 247<strong>–</strong>256 c 2009 The Authors. Journal compilation c 2009 The Zoological Society of London 255


Seasonal migration of Amazonian manatees E. M. Arraut et al.<br />

Rosas, F.C.W. (1994). Biology, conservation and status of the<br />

Amazonian manatee Trichechus inunguis. Mammal. Rev.<br />

24, 49<strong>–</strong>59.<br />

Rossetti, D.F., Toledo, P.M. & Góes, A.M. (2005). New<br />

geological framework for Western Amazonia (Brazil) and<br />

implications for biogeography and evolution. Quat. Res.<br />

63, 78<strong>–</strong>89.<br />

Schongart, J. & Junk, W.J. (2007). Forecasting the flood-pulse in<br />

Central Amazonia by ENSO-indices. J. Hydrol. 335, 124<strong>–</strong>132.<br />

Seaman, D.E. & Powell, R.A. (1996). An evaluation of the<br />

accuracy of kernel density estimators for home range<br />

analysis. Ecology. 77, 2075<strong>–</strong>2085.<br />

256<br />

Sheppard, J., Preen, A., Marsh, H., Lawler, I., Whiting, S. &<br />

Jones, R. (2006). Movement heterogeneity of dugongs,<br />

Dugong dugon (Müller), over large spatial scales. J. Exp.<br />

Mar. Biol. Ecol. 334, 64<strong>–</strong>83.<br />

Shimabukuro, Y., Novo, E.M.L.M. & Mertes, L.K. (2002).<br />

Amazonas river mainstream floodplain Landsat TM digital<br />

mosaic. Int. J. Remote Sens. 23, 57<strong>–</strong>69.<br />

Timm, R., Albuja, V. & Clauson, B. (1986). Ecology, distribution,<br />

harvest, and conservation of the Amazonian<br />

manatee Trichechus inunguis in Ecuador. Biotropica 18,<br />

150<strong>–</strong>156.<br />

Journal of Zoology 280 (2010) 247<strong>–</strong>256 c 2009 The Authors. Journal compilation c 2009 The Zoological Society of London


The Journal of Experimental Biology 212, 2349-2355<br />

Published by The Company of Biologists 2009<br />

doi:10.1242/jeb.027565<br />

Carbon and nitrogen stable isotope turnover rates and diet<strong>–</strong>tissue discrimination in<br />

Florida manatees (Trichechus manatus latirostris)<br />

Christy D. Alves-Stanley1 and Graham A. J. Worthy1,2, *<br />

1Physiological Ecology and Bioenergetics Lab, Department of Biology, University of Central Florida, 4000 Central Florida Boulevard,<br />

Orlando, FL 32816, USA and 2Hubbs-SeaWorld Research Institute, 6295 Sea Harbor Drive, Orlando, FL 32821, USA<br />

INTRODUCTION<br />

The isotopic composition of consumer tissues reflects that of local<br />

food webs and can be used to predict diet composition, the trophic<br />

level at which the consumer is feeding, and even habitat use and<br />

migratory patterns (e.g. Deniro and Epstein, 1978; Deniro and<br />

Epstein, 1981; Fry, 1981; Peterson and Fry, 1987) (reviewed by<br />

Hobson, 1999). Two stable isotope ratios commonly analyzed in<br />

feeding ecology studies are those of carbon ( 13 C/ 12 C) and nitrogen<br />

( 15 N/ 14 N). Carbon isotope ratios indicate the likely source of<br />

primary production and have been used to differentiate between C3<br />

and C4 plants, terrestrial and marine ecosystems, deep forest and<br />

open habitat consumers, and benthic and pelagic aquatic systems<br />

(e.g. Cloern et al., 2002; Cerling et al., 2004; Hall-Aspland et al.,<br />

2005). Nitrogen isotope ratios exhibit a predictable, step-wise<br />

enrichment between trophic levels and also have been shown to<br />

differ between terrestrial and marine ecosystems (e.g. Hobson and<br />

Welch, 1992).<br />

Stable isotope analysis is especially advantageous when<br />

investigating the feeding ecology and habitat use of marine mammals<br />

where it is often impossible to directly observe feeding or migratory<br />

behavior. Tissue samples, such as skin or blubber, may be analyzed<br />

for stable isotope ratios without sacrificing the animal and this<br />

approach has been successfully applied to a variety of marine<br />

mammal species including mysticetes (e.g. Lee et al., 2005),<br />

odontocetes (e.g. Walker et al., 1999), pinnipeds (e.g. Hobson and<br />

Welch, 1992; Kurle and Worthy, 2002; Newsome et al., 2006),<br />

sirenians (e.g. Ames et al., 1996; MacFadden et al., 2004; Yamamuro<br />

et al., 2004; Reich and Worthy, 2006), sea otters (Clementz and<br />

Koch, 2001) and polar bears (e.g. Ramsay and Hobson, 1991).<br />

In order to accurately interpret isotopic results, it is imperative<br />

to determine both isotopic discrimination (the difference in isotopic<br />

*Author for correspondence (e-mail: gworthy@mail.ucf.edu)<br />

Accepted 3 May 2009<br />

SUMMARY<br />

The Florida manatee (Trichechus manatus latirostris) is a herbivorous marine mammal that occupies freshwater, estuarine and<br />

marine habitats. Despite being considered endangered, relatively little is known about its feeding ecology. The present study<br />

expands on previous work on manatee feeding ecology by providing critical baseline parameters for accurate isotopic data<br />

interpretation. Stable carbon and nitrogen isotope ratios were examined over a period of more than 1 year in the epidermis of<br />

rescued Florida manatees that were transitioning from a diet of aquatic forage to terrestrial forage (lettuce). The mean half-life for<br />

13 C turnover was 53 and 59 days for skin from manatees rescued from coastal and riverine regions, respectively. The mean halflife<br />

for 15 N turnover was 27 and 58 days, respectively. Because of these slow turnover rates, carbon and nitrogen stable isotope<br />

analysis in manatee epidermis is useful in summarizing average dietary intake over a long period of time rather than assessing<br />

recent diet. In addition to turnover rate, a diet<strong>–</strong>tissue discrimination value of 2.8‰ for 13 C was calculated for long-term captive<br />

manatees on a lettuce diet. Determining both turnover rate and diet<strong>–</strong>tissue discrimination is essential in order to accurately<br />

interpret stable isotope data.<br />

Key words: turnover, stable isotope, Florida manatee, diet<strong>–</strong>tissue discrimination, 13 C, 15 N, Trichechus manatus, feeding ecology.<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY<br />

2349<br />

ratios between consumer tissue and diet) and turnover rate (the time<br />

it takes for the isotope to be assimilated into the consumer’s tissue)<br />

of the sampled tissue. Diet<strong>–</strong>tissue discrimination may be difficult<br />

to determine for animals feeding on multiple, isotopically distinct<br />

prey items for which the proportions contributing to the diet are<br />

unknown. However, controlled captive studies on a variety of taxa<br />

have allowed for more precise measurements (e.g. Roth and Hobson,<br />

2000; Cherel et al., 2005; Logan et al., 2006; Seminoff et al., 2006).<br />

In addition to diet<strong>–</strong>tissue discrimination, turnover rates in tissues<br />

must be determined in order to assess whether the isotope signature<br />

of the tissue represents the most recent diet or the long-term diet.<br />

An effective method to determine turnover rate is to experimentally<br />

switch an animal from one known diet to another isotopically distinct<br />

diet. Turnover rates of stable isotopes have been calculated using<br />

this method for mammals (e.g. Tieszen et al., 1983), birds (e.g.<br />

Hobson and Clark, 1992a), fish (e.g. Bosley et al., 2002) and<br />

invertebrates (e.g. Olive et al., 2003). These studies have shown<br />

that tissues with higher metabolic activity (e.g. blood, liver) have<br />

faster turnover rates than less active tissue (e.g. bone).<br />

The endangered Florida manatee (Trichechus manatus latirostris<br />

L.) is known to feed on a variety of aquatic plants in fresh, estuarine<br />

and marine habitats (e.g. Campbell and Irvine, 1977; Hartman, 1979;<br />

Best, 1981), each of which has a distinct isotopic signature (e.g.<br />

Fry and Sherr, 1984; Reich and Worthy, 2006) (C.D.A.-S. and<br />

G.A.J.W., in preparation). Little is known about fine-scale manatee<br />

feeding ecology and habitat use because manatees often occupy<br />

shallow, turbid water. In addition, manatee population counts and<br />

trends remain unclear (Lefebvre et al., 1995; US Fish and Wildlife<br />

Service, 2001). Consequently, it has become increasingly important<br />

to understand manatee feeding ecology and its relation to habitat<br />

use in order to improve conservation efforts.


2350<br />

C. D. Alves-Stanley and G. A. J. Worthy<br />

The present study used skin samples from Florida manatees<br />

transitioning between two isotopically distinct diets (aquatic to<br />

terrestrial) to determine turnover rates and diet<strong>–</strong>tissue discrimination<br />

in epidermis tissue. Manatees were part of the rehabilitation program<br />

at SeaWorld of Florida, and were in need of captive care for reasons<br />

including physical trauma, nutritional stress and/or cold stress. The<br />

overall objectives of the present study were to determine 13 C and<br />

15 N turnover rates in epidermis tissue and to calculate diet<strong>–</strong>tissue<br />

discrimination values for carbon and nitrogen stable isotopes in<br />

manatee skin.<br />

MATERIALS AND METHODS<br />

Sample collection<br />

Manatees held long-term at SeaWorld of Florida (Orlando, FL, USA)<br />

were fed a diet consisting primarily of romaine lettuce (>90% by<br />

mass) with minimal amounts of other terrestrial vegetation (<strong>17</strong>6 Subadult<strong>–</strong>adult 512<br />

Charlotte F 1136 330 Adult >365<br />

Primo F 494 277 Adult >365<br />

Rita F >900 >275 Adult >365<br />

Sarah F 1136 325 Adult >365<br />

Stubby F 823 252* Subadult<strong>–</strong>adult >365<br />

SWF Tm 0110 F 367 255 Subadult 1231<br />

SWF Tm 0302 F Unknown ><strong>17</strong>6 Subadult<strong>–</strong>adult 512<br />

SWF Tm 0338 F Unknown ><strong>17</strong>6 Subadult<strong>–</strong>adult 429<br />

*Missing large portion of paddle.<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY<br />

at various time intervals as they transitioned from wild forage to a<br />

captive diet. Carbon and nitrogen stable isotope turnover rates in skin<br />

were calculated for manatees rescued from two habitats in Florida:<br />

‘coastal’ (Naples on the Gulf coast and Cape Canaveral on the central<br />

east coast) and ‘riverine’ or fresh water (St Johns River near<br />

Jacksonville; Fig.1). The terms ‘coastal’ and ‘riverine’ represent rescue<br />

locations only, and are not intended to describe overall habitat use.<br />

Biopsies of epidermal tissue were collected from the trailing edge<br />

of the paddle using either a scalpel or ronguers (samples were<br />

approximately 10mm5mm, full depth of the paddle). Partially<br />

sloughed epidermis was collected directly off individual manatees<br />

(approximately 2cm2cm) if biopsies were not available. Body mass<br />

and sex were determined, and length measurements taken, where body<br />

length was measured as the straight distance from snout to paddle<br />

(O’Shea et al., 1985). Manatees were categorized into three age classes<br />

based on body length measurements (adults >275cm, subadults/late<br />

juveniles <strong>17</strong>6<strong>–</strong>275cm, and calves


Fig. 1. Rescue locations of manatees in Florida.<br />

time (days) since diet switch. Turnover rate was expressed in terms<br />

of half-life, the time it takes for the isotopic composition of the<br />

tissue to reach a midpoint between the initial and final values:<br />

X = (ln0.5) / c , (3)<br />

In order to better fit turnover data to the exponential model, an<br />

‘anchor point’ based on the mean stable isotope ratio<br />

(δ 13 C=<strong>–</strong>24.4±0.6‰, δ 15 N=2.7±0.5‰, means ± s.e.) of skin samples<br />

from nine long-term captive manatees at SeaWorld Of Florida was<br />

set at 600 days. These animals had been fed a diet of mainly lettuce<br />

for multiple years. The position of the anchor point at 600 days was<br />

chosen for several reasons: it was beyond the maximum sampling<br />

time for all rescued manatees (no manatee was sampled later than<br />

418 days), plots generally reached an asymptote at or before this<br />

point, and positions greater than 600 days did not alter results.<br />

Goodness of fit was first expressed by calculating the coefficient<br />

of determination (R 2 ) using the anchor point as part of the data set.<br />

To further illustrate fit, data for skin from each rescued manatee<br />

were paired with each individual data point contributing to the mean<br />

anchor point and minimum and maximum R 2 values were computed.<br />

All statistical analyses were judged to be significant at P


2352<br />

C. D. Alves-Stanley and G. A. J. Worthy<br />

δ 13 C (‰)<br />

δ 15 N (‰)<br />

<strong>–</strong>10<br />

<strong>–</strong>12<br />

<strong>–</strong>14<br />

<strong>–</strong>16<br />

<strong>–</strong>18<br />

<strong>–</strong>20<br />

<strong>–</strong>22<br />

<strong>–</strong>24<br />

<strong>–</strong>26<br />

<strong>–</strong>28<br />

11<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

0 100 200 300 400 500 600<br />

1<br />

0 100 200 300 400 500 600<br />

manatees were also fitted to the exponential decay model (Fig.2).<br />

Carbon turnover half-lives in the skin of riverine manatees ranged<br />

from 39 to 72 days with a mean of 59±18 days; however, a halflife<br />

was not calculated for manatee 0341 because the equation<br />

showed little to no change in signature over time (Fig.2; Table4).<br />

Skin samples from these riverine manatees were also enriched<br />

in 15 N relative to that of captive manatees (mean enrichment=<br />

6.3±1.6‰) and nitrogen half-lives ranged from 21 to 115 days with<br />

a mean of 58±42 days (Fig.2; Table4). These turnover times were<br />

not significantly different from carbon half-lives (paired t-test,<br />

t=0.13, d.f.=2, P=0.91). MANOVA results indicated there were no<br />

significant differences in stable carbon or nitrogen isotope half-lives<br />

between manatees rescued from riverine vs coastal regions (F-test:<br />

F2,4=0.58, P=0.60).<br />

DISCUSSION<br />

Carbon enrichment values calculated in the present study<br />

(2.8±0.9‰) were similar to values previously reported. Ames and<br />

colleagues (Ames et al., 1996) found sloughed skin from captive<br />

manatees to be enriched in 13 C by an average of 4.1‰ compared<br />

with lettuce. Reich and Worthy (Reich and Worthy, 2006) assumed<br />

a carbon enrichment value of 3.0‰ in manatee skin when applying<br />

an isotope mixing model (Phillips and Gregg, 2001) to diet<br />

interpretation of free-ranging manatees. The only other known study<br />

on diet<strong>–</strong>tissue discrimination in any mammalian skin is that of<br />

Hobson and colleagues (Hobson et al., 1996) in which seal skin<br />

was found to be enriched in 13 C relative to diet by 2.8‰. In the<br />

present study, nitrogen enrichment could not be determined because<br />

Coastal manatees Riverine manatees<br />

Animal ID Half-life<br />

<strong>–</strong>10<br />

Animal ID Half-life<br />

0301 63 days <strong>–</strong>12<br />

0334 72 days<br />

0318 60 days<br />

0322 45 days<br />

<strong>–</strong>14<br />

0340 39 days<br />

0341 N/A<br />

0431 42 days <strong>–</strong>16<br />

0501 67 days<br />

Animal ID Half-life<br />

0301 36 days<br />

0318 33 days<br />

0322 14 days<br />

0431 23 days<br />

<strong>–</strong>18<br />

<strong>–</strong>20<br />

<strong>–</strong>22<br />

<strong>–</strong>24<br />

<strong>–</strong>26<br />

<strong>–</strong>28<br />

11<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0 100 200 300 400 500 600<br />

Days since diet change<br />

0 100 200 300 400 500 600<br />

Animal ID Half-life<br />

0334 115 days<br />

0340 34 days<br />

0341 62 days<br />

0501 21 days<br />

Fig. 2. 13 C and 15 N turnover in epidermis from manatees rescued near Naples and Cape Canaveral (coastal manatees) and from the St Johns River (riverine<br />

manatees) in Florida. Mean (±95% CI) stable isotope ratios for skin from long-term captive manatees () were used as ‘anchor points’ set at 600 days to<br />

better fit the models. To illustrate diet<strong>–</strong>tissue discrimination, mean (±95% CI) stable isotope ratios for the main diet items fed in captivity (romaine lettuce and<br />

spinach) are indicated by a horizontal solid black line and horizontal dashed lines.<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY<br />

of the high degree of variability of nitrogen signatures in the lettuce<br />

diet. Typically, diet<strong>–</strong>tissue discrimination values for nitrogen are in<br />

the range 2<strong>–</strong>5‰ (Peterson and Fry, 1987; Kelly, 2000).<br />

Manatees rescued from coastal regions were ideal subjects for<br />

carbon isotope turnover calculations because carbon signatures in<br />

their skin differed dramatically from those of captive manatees.<br />

Interpreting δ 13 C values in the skin of riverine manatees was<br />

problematic because of variability in values at the time of rescue<br />

and the similarity of δ 13 C values between the skin of rescued<br />

manatees and that of long-term captive manatees. The half-lives in<br />

skin from riverine manatees were not significantly different from<br />

those of skin from coastal manatees; however, the carbon turnover<br />

data for manatees rescued from riverine regions did not fit the<br />

exponential decay models as closely as those for coastal manatees.<br />

The carbon half-life calculated for manatee epidermis was very<br />

slow compared with previous turnover studies on other species.<br />

Stable isotope turnover rates can differ based on the particular<br />

isotope, tissue and/or taxon analyzed, diet, physiological state,<br />

feeding rate and/or growth rate of the animal (e.g. Fry and Arnold,<br />

1982; Bosley et al., 2002; Hobson and Bairlein, 2003; Olive et al.,<br />

2003). Additionally, some studies removed lipids from samples<br />

while others did not. Therefore, direct comparisons between studies<br />

are difficult. Dalerum and Angerbjorn (Dalerum and Angerbjorn,<br />

2005) cautioned that comparisons of turnover rates should be made<br />

between the same tissues to avoid these complications. Additionally,<br />

turnover rates in tissues should be compared between animals of<br />

similar body size as metabolic rates have an effect on isotope<br />

turnover (Sponheimer et al., 2006).


The present study is the first to calculate stable isotope turnover<br />

rates in the skin of any marine mammal species. Isotope turnover<br />

rates that have been reported for large terrestrial mammals including<br />

bears (Hilderbrand et al., 1996), alpacas (Sponheimer et al., 2006),<br />

and domestic cattle and horses (Schwertl et al., 2003; Ayliffe et al.,<br />

2004) were determined using blood, muscle and liver, and hair,<br />

respectively. As no appropriate comparison between turnover rates<br />

in the skin of large mammals was possible, the results from this<br />

study will be cautiously compared with others. Previous studies on<br />

stable isotope ratios in hair were omitted from this comparison as<br />

hair is a metabolically inert tissue in which the isotopic composition<br />

represents the period of growth.<br />

The only reported carbon half-lives in animal tissue that are<br />

greater than those of manatee epidermis are those of alpaca muscle<br />

[<strong>17</strong>9 days (Sponheimer et al., 2006)], bat wing membrane and whole<br />

blood [102<strong>–</strong>134 days (Voigt et al., 2003)], and quail bone collagen<br />

[<strong>17</strong>3 days (Hobson and Clark, 1992a)]. In the alpaca study, muscle<br />

tissues were not lipid extracted so direct comparisons may be<br />

problematic. Voigt and colleagues (Voigt et al., 2003) suggested<br />

the slow turnover rate in bat wing membrane was largely due to the<br />

tissue being composed primarily of collagen and elastin, which are<br />

known to have slow turnover rates in bone. Additionally, Voigt and<br />

colleagues (Voigt et al., 2003) attributed the slow turnover rate in<br />

bat blood to their long-lived erythrocytes. Finally, the long half-life<br />

in quail bone aligns with collagen being a less metabolically active<br />

tissue. Epidermal tissue is composed of keratin in the epithelial<br />

lamina and collagen and elastin in the basal lamina. Manatee<br />

epidermis has been described as thick and possesses the<br />

characteristic of hyperkeratosis (Sokolov, 1982; Graham et al.,<br />

2003). It is possible that a slow replacement of keratin in manatee<br />

Manatee stable isotope turnover<br />

Table 4. Exponential decay equations and half-lives representing stable isotope turnover in epidermis sampled from rehabilitated Florida<br />

manatees<br />

R 2 range<br />

Animal Half-life δ13C at<br />

Carbon turnover ID Equation R2 (days) day 0 (‰) Min. Max.<br />

Coastal manatees 0301 y=<strong>–</strong>24.4+14.2e <strong>–</strong>0.01094x 1.00 63 <strong>–</strong>10.2 1.00 1.00<br />

0318 y=<strong>–</strong>24.8+15.1e <strong>–</strong>0.01158x 0.97 60 <strong>–</strong>9.7 0.96 0.97<br />

0322 y=<strong>–</strong>24.5+14.3e <strong>–</strong>0.01550x 0.83 45 <strong>–</strong>10.2 0.79 0.84<br />

0431 y=<strong>–</strong>24.1+11.6e <strong>–</strong>0.01657x 0.97 42 <strong>–</strong>12.5 0.97 0.97<br />

Mean 53 <strong>–</strong>10.7<br />

Riverine manatees 0334 y=<strong>–</strong>24.6+8.2e <strong>–</strong>0.00963x 0.95 72 <strong>–</strong>16.4 0.90 0.96<br />

0340 y=<strong>–</strong>24.7+0.9e <strong>–</strong>0.0<strong>17</strong>87x 0.36 39 <strong>–</strong>23.8 0.05 0.51<br />

0341 y=<strong>–</strong>23.8+1.9e <strong>–</strong>20920x 0.50 n.d. <strong>–</strong>21.9 0.43 0.51<br />

0501 y=<strong>–</strong>24.5+6.2e <strong>–</strong>0.01028x 0.70 67 <strong>–</strong>18.3 0.59 0.75<br />

Mean 59 <strong>–</strong>20.1<br />

R 2 range<br />

Animal Half-life δ15N at<br />

Nitrogen turnover ID Equation R2 (days) day 0 (‰) Min. Max.<br />

Coastal manatees 0301 y=2.3+3.4e <strong>–</strong>0.01918x 0.96 36 5.7 0.83 1.00<br />

0318 y=2.4+2.4e <strong>–</strong>0.02125x 0.57 33 4.8 0.46 0.62<br />

0322 y=2.3+3.6e <strong>–</strong>0.04898x 0.56 14 5.9 0.29 0.68<br />

0431 y=2.5+4.7e <strong>–</strong>0.03038x 0.97 23 7.2 0.91 0.97<br />

Mean 27 5.9<br />

Riverine manatees 0334 y=2.6+4.9e <strong>–</strong>0.00601x 1.00 115 7.5 0.98 1.00<br />

0340 y=2.7+6.4e <strong>–</strong>0.02055x 0.92 34 9.1 0.90 0.92<br />

0341 y=2.3+5.9e <strong>–</strong>0.01125x 0.85 62 8.2 0.80 0.87<br />

0501 y=2.0+9.4e <strong>–</strong>0.03273x 0.95 21 11.4 0.92 0.97<br />

n.d., not determined (see text).<br />

Mean 58 9.1<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY<br />

2353<br />

epidermis and the presence of collagen and elastin in the basal lamina<br />

also contributed to the slow isotope turnover rate in the skin.<br />

Another potentially significant factor impacting on turnover rate<br />

is the manatees’ overall slow metabolism. Metabolic rates in adult<br />

Florida manatees have been shown to be lower than those predicted<br />

based on body size [15<strong>–</strong>40% of predicted values (Irvine, 1983;<br />

Worthy et al., 2000)]. It is also possible that food passage time<br />

impacts on turnover rate [as discussed by Post (Post, 2002)].<br />

Manatees use hindgut fermentation and have a passage rate through<br />

the digestive tract of 146<strong>–</strong>147h (Lomolino and Ewel, 1984; Larkin<br />

et al., 2007). This rate is consistent with that of the dugong, another<br />

sirenian [145<strong>–</strong>166h (Lanyon and Marsh, 1995)], but much slower<br />

than those of other large hindgut fermenters such as elephants<br />

[21<strong>–</strong>46h (Rees, 1982)], horses [26<strong>–</strong>27h (Rosenfeld et al., 2006)]<br />

and rhinoceros [61h (Clauss et al., 2005)]. Manatees sampled in<br />

the present study were rescued for reasons including cold stress,<br />

entanglement and watercraft injuries. Because of their physical<br />

condition, their intake rates may have been slower than those of<br />

manatees not in need of rehabilitation.<br />

The gruel mixture fed to the manatees for the first few weeks of<br />

rehabilitation was enriched in 13 C compared with lettuce and<br />

spinach. It is presumed that the initial supplementation of the diet<br />

with gruel would have had minimal to no impact on the carbon<br />

turnover rate as turnover was already very slow. There were no<br />

significant differences in δ 13 C values between biopsy and sloughed<br />

skin samples, so the differing sample types had no effect on carbon<br />

turnover rate. Carbon isotope ratios of manatee fecal material did<br />

not differ from those of the main diet items, so even if manatees<br />

were engaging in coprophagy it would not have had any effect on<br />

carbon turnover rate in the skin. This result is another indication


2354<br />

C. D. Alves-Stanley and G. A. J. Worthy<br />

that stable isotope analysis of fecal material has great potential in<br />

assessing short-term, recent dietary choices.<br />

Phillips and Koch (Phillips and Koch, 2002) suggested<br />

incorporating carbon and nitrogen concentration analysis to aid in<br />

stable isotope dietary reconstruction, especially when there are great<br />

differences in C and N concentration between diet items. We were<br />

unable to include a concentration analysis in this study; however,<br />

it may be a useful addition to future analyses.<br />

Coastal manatees were poor subjects for nitrogen turnover<br />

calculations because of the high variability in their initial δ 15 N values<br />

at the time of rescue, the similarity of δ 15 N values between the skin<br />

of rescued animals and that of captive animals, and the high<br />

variability in δ 15 N values of captive diet items. Consequently,<br />

calculated nitrogen half-lives were inconsistent. Riverine manatees<br />

were better subjects because there was a greater difference in<br />

nitrogen signatures of skin between rescued and long-term captive<br />

animals. Even then, nitrogen half-lives for skin from riverine<br />

manatees were still variable, most likely because of the variability<br />

in nitrogen signatures of romaine lettuce and spinach fed in captivity.<br />

The lettuce and spinach in the captive manatee diet often originated<br />

from different agricultural producers and it is possible that different<br />

fertilization techniques were used resulting in high variability in<br />

δ 15 N values (see Georgi et al., 2005).<br />

There are very few studies on nitrogen turnover in other species.<br />

Nitrogen half-lives in mammalian tissues have been calculated for<br />

blood plasma and cells in black bears [3 and 22 days, respectively<br />

(Hilderbrand et al., 1996)] and avian whole blood [10.0<strong>–</strong>14.4 days<br />

(Bearhop et al., 2002; Hobson and Bairlein, 2003; Ogden et al.,<br />

2004)] and plasma [0.5<strong>–</strong>1.7 days (Pearson et al., 2003)]. Nitrogen<br />

turnover in manatee skin was relatively slow compared with the<br />

results of these studies, most likely owing to the low metabolic rate<br />

of manatees, epidermal tissue composition and/or food passage rate<br />

as previously discussed in terms of carbon turnover.<br />

There was no compounding effect of the gruel supplement on<br />

nitrogen turnover rate as the signature of the gruel was not<br />

significantly different from that of romaine lettuce and spinach.<br />

Likewise, coprophagy would have had no effect on nitrogen<br />

turnover rate in manatee skin as the δ 15 N values of manatee fecal<br />

material did not differ from those of the main diet items. However,<br />

sloughed skin samples were significantly enriched in 15 N compared<br />

with biopsy samples. Though δ 15 N values were adjusted to account<br />

for this enrichment, variability in nitrogen turnover rates and lack<br />

of fit indicate that sample type may have contributed to the difficulty<br />

in calculating more precise half-lives. At the present time, it is<br />

unclear why sloughed samples differed in δ 15 N values, but not δ 13 C<br />

values, from biopsy samples. It is possible that materials were<br />

redistributed within the skin while sloughing, and sloughed samples<br />

also contained less epidermal depth than biopsies. Both of these<br />

factors could have contributed to differing δ 15 N values between<br />

sample types.<br />

Cerling and colleagues (Cerling et al., 2007) and Ayliffe and<br />

colleagues (Ayliffe et al., 2004) have suggested that some stable<br />

isotope turnover data may be better fitted to a multiple-pool model<br />

than to the traditional single-pool, exponential decay model used<br />

in the present study. Proper application requires knowledge of each<br />

pool in vivo (Sponheimer et al., 2006), which at this time is unknown<br />

for manatees. Consequently, we did not incorporate a multiple-pool<br />

analysis. Single-pool analyses are still useful for intraspecific and<br />

interspecific comparisons of stable isotope turnover (Sponheimer<br />

et al., 2006).<br />

When proportions of food sources contributing to a mixed diet<br />

are unknown, mixing models are often used to aid in estimating these<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY<br />

proportions (e.g. Phillips and Gregg, 2001; Newsome et al., 2004;<br />

Reich and Worthy, 2006). If a change in diet occurs, the resulting<br />

signature may not be representative of the current diet but, in fact,<br />

may be some intermediate value between two distinct diets. While<br />

this possibility is true for all stable isotope analyses, the impact is<br />

minimized in tissues with high turnover rates because the time frame<br />

is short. Free-ranging manatees are known to switch diet sources<br />

(Best, 1981; Lefebvre et al., 2000), and the very slow turnover rates<br />

for carbon and nitrogen stable isotopes in epidermis tissue mean that<br />

unless the manatee has been feeding on the same diet for an extended<br />

period of time, the skin signature will always be in some transitional<br />

state. Slow turnover rates in manatee skin especially complicate<br />

estimation of the proportions of freshwater, estuarine and marine<br />

sources in the diet because δ 13 C values for estuarine vegetation are<br />

intermediate between those of freshwater vegetation and seagrasses<br />

(Reich and Worthy, 2006) (C.D.A.-S. and G.A.J.W., in preparation).<br />

The incorporation of nitrogen stable isotope analysis can aid in further<br />

separation of these three diet sources as nitrogen signatures for<br />

freshwater and estuarine vegetation differ from those of seagrasses<br />

(C.D.A.-S. and G.A.J.W., in preparation).<br />

Computing a precise diet<strong>–</strong>tissue discrimination value is essential<br />

when interpreting isotopic results. Discrimination values for carbon<br />

have previously been calculated for manatee skin on a captive diet<br />

(Ames et al., 1996) and discrimination values for carbon and nitrogen<br />

have been estimated in the skin of free-ranging manatees on<br />

possible diets of freshwater, estuarine and/or marine vegetation<br />

(Reich and Worthy, 2006). It is unknown whether diet<strong>–</strong>tissue<br />

discrimination in manatee skin differs between diet types as has<br />

been shown in other studies (e.g. Hobson and Clark, 1992b). The<br />

results of the present study are the most extensive thus far.<br />

Carbon and nitrogen stable isotope analysis of manatee epidermal<br />

tissue is difficult, if not impossible, to use when assessing shortterm<br />

or recent changes in diet and habitat use because of slow<br />

turnover rates. This technique would potentially have more direct<br />

application in summarizing average dietary intake over longer<br />

periods of time. In order to accurately interpret isotopic analyses,<br />

determining diet<strong>–</strong>tissue discrimination factors and turnover rates in<br />

the tissue is essential. The difficulty with most studies is that isotope<br />

discrimination and turnover are best calculated under controlled<br />

situations in captivity. Mixing model results for tissues with slow<br />

turnover rates should be interpreted with caution, especially in<br />

species that may be switching between diets in which an intermediate<br />

isotope ratio may be mistakenly described as indicating a single<br />

diet source instead of a mixture of sources.<br />

We thank R. Bonde, M. Ross and B. Chittick for assistance with manatee skin<br />

sample collections and we are grateful for funding from the University of Central<br />

Florida (UCF) Department of Biology and Graduate Studies to C.D.A.-S. and a<br />

Provosts Research Enhancement Award to G.A.J.W. Manatee research was<br />

carried out under UCF-Institutional Animal Care and Use Committee protocol 02-<br />

09W and US Fish and Wildlife Service (USFWS) permit number MA056326<br />

issued to G.A.J.W. and USFWS permit number MA79<strong>17</strong>21 issued to R. Bonde.<br />

Plant samples were collected under Florida Department of Environmental<br />

Protection plant collection permit number <strong>17</strong>53 issued to G.A.J.W. We thank J.<br />

Roth and J. Weishampel for assistance with editing and J. Fauth for assistance<br />

with statistical analyses. Additional thanks to T. Maddox, R. Harris, R. Runnels, T.<br />

Doyle, K. Fuhr, J. Greenawalt, L. Hoopes, A. Stephens, M. DiPiazza, N. Browning<br />

and J. Stanley for field and lab assistance.<br />

References<br />

Ames, A. L., VanVleet, E. S. and Sackett, W. M. (1996). The use of stable carbon<br />

isotope analysis for determining the dietary habits of the Florida manatee,<br />

Trichechus manatus latirostris. Mar. Mamm. Sci. 12, 555-563.<br />

Ayliffe, L. K., Cerling, T. E., Robinson, T., West, A. G., Sponheimer, M., Passey,<br />

B. H., Hammer, J., Roeder, B., Dearing, M. D. and Ehleringer, J. R. (2004).<br />

Turnover of carbon isotopes in tail hair and breath CO2 of horses fed an isotopically<br />

varied diet. Oecologia 139, 11-22.


Bearhop, S., Waldron, S., Votier, S. C. and Furness, R. W. (2002). Factors that<br />

influence assimilation rates and fractionation of nitrogen and carbon stable isotopes<br />

in avian blood and feathers. Physiol. Biochem. Zool. 75, 451-458.<br />

Best, R. C. (1981). Foods and feeding habits of wild and captive Sirenia. Mamm. Rev.<br />

11, 3-29.<br />

Bosley, K. L., Witting, D. A., Chambers, R. C. and Wainright, S. C. (2002).<br />

Estimating turnover rates of carbon and nitrogen in recently metamorphosed winter<br />

flounder Pseudopleuronectes americanus with stable isotopes. Mar. Ecol. Prog. Ser.<br />

236, 233-240.<br />

Campbell, H. W. and Irvine, A. B. (1977). Feeding ecology of West Indian manatee<br />

Trichechus manatus Linnaeus. Aquaculture 12, 249-251.<br />

Cerling, T. E., Hart, J. A. and Hart, T. B. (2004). Stable isotope ecology in the Ituri<br />

Forest. Oecologia 138, 5-12.<br />

Cerling, T. E., Ayliffe, L. K., Dearing, M. D., Ehleringer, J. R., Passey, B. H.,<br />

Podlesak, D. W. A., T. and West, A. G. (2007). Determining biological tissue turnover<br />

using stable isotopes: the reaction progress variable. Oecologia 151, <strong>17</strong>5-189.<br />

Cherel, Y., Hobson, K. A. and Hassani, S. (2005). Isotopic discrimination between<br />

food and blood and feathers of captive penguins: implications for dietary studies in<br />

the wild. Physiol. Biochem. Zool. 78, 106-115.<br />

Clauss, M., Polster, C., Kienzle, E., Wiesner, H., Baumgartner, K., von Houwald,<br />

F., Ortmann, S., Streich, W. J. and Dierenfeld, E. S. (2005). Studies on digestive<br />

physiology and feed digestibilities in captive Indian rhinoceros (Rhinoceros<br />

unicornis). J. Anim. Physiol. Anim. Nutr. 89, 229-237.<br />

Clementz, M. T. and Koch, P. L. (2001). Differentiating aquatic mammal habitat and<br />

foraging ecology with stable isotopes in tooth enamel. Oecologia 129, 461-472.<br />

Cloern, J. E., Canuel, E. A. and Harris, D. (2002). Stable carbon and nitrogen<br />

isotope composition of aquatic and terrestrial plants of the San Francisco Bay<br />

estuarine system. Limnol. Oceanogr. 47, 713-729.<br />

Craig, H. (1957). Isotopic standards for hydrogen and oxygen and correlation factors<br />

for mass-spectrometric analysis of carbon dioxide. Geochim. Cosmochim. Acta 42,<br />

495-506.<br />

Dalerum, F. and Angerbjorn, A. (2005). Resolving temporal variation in vertebrate<br />

diets using naturally occurring stable isotopes. Oecologia 144, 647-658.<br />

Deniro, M. J. and Epstein, S. (1978). Influence of diet on distribution of carbon<br />

isotopes in animals. Geochim. Cosmochim. Acta 42, 495-506.<br />

Deniro, M. J. and Epstein, S. (1981). Influence of diet on the distribution of nitrogen<br />

isotopes in animals. Geochim. Cosmochim. Acta 45, 341-351.<br />

Fry, B. (1981). Natural stable carbon isotope tag traces Texas shrimp migrations. Fish.<br />

Bull. 79, 337-345.<br />

Fry, B. and Arnold, C. (1982). Rapid C-13/C-12 turnover during growth of brown<br />

shrimp (Penaeus aztecus). Oecologia 54, 200-204.<br />

Fry, B. and Sherr, E. B. (1984). δ 13 C measurements as indicators of carbon flow in<br />

marine and freshwater ecosystems. Contrib. Mar. Sci. 27, 13-47.<br />

Georgi, M., Voerkelius, S., Rossmann, A., Grassmann, J. and Schnitzler, W. H.<br />

(2005). Multielement isotope ratios of vegetables from integrated and organic<br />

production. Plant Soil 275, 93-100.<br />

Graham, A. R., Samuelson, D. A., Isaza, R. and Lewis, P. A. (2003). Histological<br />

comparison of manatee and elephant integument. In 15th Biennial Conference on<br />

the Biology of Marine Mammals, pp. 62-63. Greensboro, NC: Society for Marine<br />

Mammalogy.<br />

Hall-Aspland, S. A., Rogers, T. L. and Canfield, R. B. (2005). Stable carbon and<br />

nitrogen isotope analysis reveals seasonal variation in the diet of leopard seals. Mar.<br />

Ecol. Prog. Ser. 305, 249-259.<br />

Hartman, D. S. (1979). Ecology and behavior of the manatee (Trichechus manatus) in<br />

Florida: The American Society of Mammalogists Special Publication No. 5.<br />

Hilderbrand, G. V., Farley, S. D., Robbins, C. T., Hanley, T. A., Titus, K. and<br />

Servheen, C. (1996). Use of stable isotopes to determine diets of living and extinct<br />

bears. Can. J. Zool. 74, 2080-2088.<br />

Hobson, K. A. (1999). Tracing origins and migration of wildlife using stable isotopes: a<br />

review. Oecologia 120, 314-326.<br />

Hobson, K. A. and Bairlein, F. (2003). Isotopic fractionation and turnover in captive<br />

garden warblers (Sylvia borin): implications for delineating dietary and migratory<br />

associations in wild passerines. Can. J. Zool. 81, 1630-1635.<br />

Hobson, K. A. and Clark, R. G. (1992a). Assessing avian diets using stable isotopes<br />

1. Turnover of 13 C in tissues. Condor 94, 181-188.<br />

Hobson, K. A. and Clark, R. G. (1992b). Assessing avian diets using stable isotopes<br />

2. Factors influencing diet-tissue fractionation. Condor 94, 189-197.<br />

Hobson, K. A. and Welch, H. E. (1992). Determination of trophic relationships within<br />

a high Arctic marine food web using δ 13 C and δ 15 N analysis. Mar. Ecol. Prog. Ser.<br />

84, 9-18.<br />

Hobson, K. A., Schell, D. M., Renouf, D. and Noseworthy, E. (1996). Stable carbon<br />

and nitrogen isotopic fractionation between diet and tissues of captive seals:<br />

implications for dietary reconstructions involving marine mammals. Can. J. Fish<br />

Aquat. Sci. 53, 528-533.<br />

Irvine, A. B. (1983). Manatee metabolism and its influence on distribution in Florida.<br />

Biol. Conserv. 25, 315-334.<br />

Kelly, J. F. (2000). Stable isotopes of carbon and nitrogen in the study of avian and<br />

mammalian trophic ecology. Can. J. Zool. 78, 1-27.<br />

Kurle, C. M. and Worthy, G. A. J. (2002). Stable nitrogen and carbon isotope ratios in<br />

multiple tissues of the northern fur seal Callorhinus ursinus: implications for dietary<br />

and migratory reconstructions. Mar. Ecol. Prog. Ser. 236, 289-300.<br />

Lanyon, J. M. and Marsh, H. (1995). Digesta passage times in the dugong. Aust. J.<br />

Zool. 43, 119-127.<br />

Larkin, I. L. V., Fowler, V. F. and Reep, R. L. (2007). Digesta passage rates in the<br />

Florida manatee (Trichechus manatus latirostris). Zoo Biol. 26, 503-515.<br />

Lee, S. H., Schell, D. M., McDonald, T. L. and Richardson, W. J. (2005). Regional<br />

and seasonal feeding by bowhead whales Balaena mysticetus as indicated by stable<br />

isotope ratios. Mar. Ecol. Prog. Ser. 285, 271-287.<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY<br />

Manatee stable isotope turnover<br />

2355<br />

Lefebvre, L. W., Ackerman, B. B., Portier, K. M. and Pollock, K. H. (1995). Aerial<br />

survey as a technique for estimating trends in manatee population size: problems<br />

and prospects. In Population biology of the Florida Manatee (ed. T. J. O’Shea, B. B.<br />

Ackerman and H. F. Percival), pp. 63-74. Washington, DC: Information and<br />

Technology Report 1, National Biological Service.<br />

Lefebvre, L. W., Reid, J. P., Kenworthy, W. J. and Powell, J. A. (2000).<br />

Characterizing manatee habitat use and seagrass grazing in Florida and Puerto<br />

Rico: implications for conservation and management. Pac. Conserv. Biol. 5, 289-<br />

298.<br />

Logan, J., Haas, H., Deegan, L. and Gaines, E. (2006). Turnover rates of nitrogen<br />

stable isotopes in the salt marsh mummichog, Fundulus heteroclitus, following a<br />

laboratory diet switch. Oecologia 147, 391-395.<br />

Lomolino, M. W. and Ewel, K. C. (1984). Digestive efficiencies of the West Indian<br />

manatee (Trichechus manatus). Florida Sci. 47, <strong>17</strong>6-<strong>17</strong>9.<br />

MacFadden, B. J., Higgins, P., Clementz, M. T. and Jones, D. S. (2004). Diets,<br />

habitat preferences, and niche differentiation of Cenozoic sirenians from Florida:<br />

evidence from stable isotopes. Paleobiology 30, 297-324.<br />

Newsome, S. D., Phillips, D. L., Culleton, B. J., Guilderson, T. P. and Koch, P. L.<br />

(2004). Dietary reconstruction of an early to middle Holocene human population from<br />

the central California coast: insights from advanced stable isotope mixing models. J.<br />

Archaeol. Sci. 31, 1101-1115.<br />

Newsome, S. D., Koch, P. L., Etnier, M. A. and Aurioles-Gambao, D. (2006). Using<br />

carbon and nitrogen isotope values to investigate maternal strategies in northeast<br />

Pacific otariids. Mar. Mamm. Sci. 22, 556-572.<br />

O’Shea, T. J., Beck, C. A., Bonde, R. K., Kochman, H. I. and Odell, D. K. (1985).<br />

An analysis of manatee mortality patterns in Florida, 1976-81. J. Wildl. Manage. 49,<br />

1-11.<br />

Ogden, L. J. E., Hobson, K. A. and Lank, D. B. (2004). Blood isotopic (d 13 C and<br />

d 15 N) turnover and diet-tissue fractionation factors in captive Dunlin (Calidris alpina<br />

pacifica). Auk 121, <strong>17</strong>0-<strong>17</strong>7.<br />

Olive, P. J. W., Pinnegar, J. K., Polunin, N. V. C., Richards, G. and Welch, R.<br />

(2003). Isotope trophic-step fractionation: a dynamic equilibrium model. J. Anim.<br />

Ecol. 72, 608-6<strong>17</strong>.<br />

Pearson, S. F., Levey, D. J., Greenberg, C. H. and del Rio, C. M. (2003). Effects of<br />

elemental composition on the incorporation of dietary nitrogen and carbon isotopic<br />

signatures in an omnivorous songbird. Oecologia 135, 516-523.<br />

Peterson, B. J. and Fry, B. (1987). Stable isotopes in ecosystem studies. Annu. Rev.<br />

Ecol. Syst. 18, 293-320.<br />

Phillips, D. L. and Gregg, J. W. (2001). Uncertainty in source partitioning using stable<br />

isotopes. Oecologia 127, <strong>17</strong>1-<strong>17</strong>9.<br />

Phillips, D. L. and Koch, P. L. (2002). Incorporating concentration dependence in<br />

stable isotope mixing models. Oecologia 130, 114-125.<br />

Post, D. M. (2002). Using stable isotopes to estimate trophic position: models,<br />

methods, and assumptions. Ecology 83, 703-718.<br />

Ramsay, M. A. and Hobson, K. A. (1991). Polar bears make little use of terrestrial<br />

food webs: evidence from stable-carbon isotope analysis. Oecologia 86, 598-600.<br />

Rau, G. H., Ainley, D. G., Bengtson, J. L., Torres, J. J. and Hopkins, T. L. (1992).<br />

15 N/ 14 N and 13 C/ 12 C in Weddell Sea birds, seals, and fish: implications for diet and<br />

trophic structure. Mar. Ecol. Prog. Ser. 84, 1-8.<br />

Rees, R. A. (1982). Gross assimilation efficiency and food passage time in the African<br />

elephant. Afr. J. Ecol. 20, 193-198.<br />

Reich, K. J. and Worthy, G. A. J. (2006). An isotopic assessment of the feeding<br />

habits of free-ranging manatees. Mar. Ecol. Prog. Ser. 322, 303-309.<br />

Rosenfeld, I., Austbo, D. and Volden, H. (2006). Models for estimating digesta<br />

passage kinetics in the gastrointestinal tract of the horse. J. Anim. Sci. 83, 3321-<br />

3328.<br />

Roth, J. D. and Hobson, K. A. (2000). Stable carbon and nitrogen isotopic<br />

fractionation between diet and tissue of captive red fox: implications for dietary<br />

reconstruction. Can. J. Zool. 78, 848-852.<br />

Schwertl, M., Auerswald, K. and Schnyder, H. (2003). Reconstruction of the isotopic<br />

history of animal diets by hair segmental analysis. Rapid Commun. Mass Spectrom.<br />

<strong>17</strong>, 1312-1318.<br />

Seminoff, J. A., Jones, T. T., Eguchi, T., Jones, D. R. and Dutton, P. H. (2006).<br />

Stable isotope discrimination (δ 13 C and δ 15 N) between soft tissues of the green sea<br />

turtle Chelonia mydas and its diet. Mar. Ecol. Prog. Ser. 308, 271-278.<br />

Sokolov, V. E. (1982). Mammal Skin. Berkeley, CA: University of California Press.<br />

Sponheimer, M., Robinson, T. F., Cerling, T. E., Tegland, L., Roeder, B. L., Ayliffe,<br />

L., Dearing, M. D. and Ehleringer, J. R. (2006). Turnover of stable carbon isotopes<br />

in the muscle, liver, and breath CO2 of alpacas (Lama pacos). Rapid Commun.<br />

Mass Spectrom. 20, 1395-1399.<br />

Tieszen, L. L., Boutton, T. W., Tesdahl, K. G. and Slade, N. A. (1983). Fractionation<br />

and turnover of stable carbon isotopes in animal tissues: implications for δ 13 C<br />

analysis of diet. Oecologia 57, 32-37.<br />

US Fish and Wildlife Service (2001). Florida Manatee Recovery Plan (Trichechus<br />

manatus latirostris), 3rd revision. Atlanta, GA: US Fish and Wildlife Service.<br />

Voigt, C. C., Matt, F., Michener, R. and Kunz, T. H. (2003). Low turnover rates of<br />

carbon isotopes in tissues of two nectar-feeding bat species. J. Exp. Biol. 206, 1419-<br />

1427.<br />

Walker, J. L., Potter, C. W. and Macko, S. A. (1999). The diets of modern and<br />

historic bottlenose dolphin populations reflected through stable isotopes. Mar.<br />

Mamm. Sci. 15, 335-350.<br />

Worthy, G. A. J., Miculka, T. A. and Wright, S. D. (2000). Manatee response to cold:<br />

how cold is too cold? In Florida Manatees and Warm Water: Proceedings of the<br />

Warm Water Workshop (ed. W. Perkins), pp. 1-6. Jacksonville, FL: US Fish and<br />

Wildlife Service.<br />

Yamamuro, M., Aketa, K. and Uchida, S. (2004). Carbon and nitrogen stable isotope<br />

ratios of the tissues and gut contents of a dugong from the temperate coast of<br />

Japan. Mammal Study 29, <strong>17</strong>9-183.


Trend Detection in a Boat-based Method for Monitoring <strong>Sirenian</strong>s: Antillean Manatee<br />

Case Study<br />

Katherine S. LaCommare a* , Solange Brault a , Caryn Self-Sullivan b , Ellen M. Hines c<br />

a University of Massachusetts, Boston 100 Morrissey Blvd. Boston, MA 02125<br />

katie.lacommare@umb.edu, solange.brault@umb.edu<br />

b <strong>Sirenian</strong> <strong>International</strong>, Inc., 200 Stonewall Dr. Fredericksburg, VA 22401 caryns@sirenian.org<br />

c Department of Geography and Human Environmental Studies, San Francisco State University,<br />

1600 Holloway Ave. San Francisco, CA 94132 ehines@sfsu.edu<br />

Abstract.
Accurate monitoring is a critical step in evaluating the conservation and management<br />

needs of endangered species. We evaluated a low cost, effective survey method for monitoring<br />

West Indian manatees (Trichechus manatus manatus) in Belize, Central America. The<br />

objectives for this paper are (1) to evaluate a count-based population index derived from a boatbased<br />

survey method, (2) to examine trends in manatee abundance in the Drowned Cayes area,<br />

(3) to conduct a power analysis to explore our ability to detect a trend and the ramifications of<br />

survey structure on trend detection. We used a generalized linear model to evaluate the impact<br />

of environmental conditions on sighting probability and to determine whether the number of<br />

manatees observed per 20-minute scan changed from 2001 - 2007. We used simulations to<br />

determine statistical power - the ability to detect potential declines of 10%, 25% or 50% over 15<br />

years and for various sampling regimes. The number of manatees sighted per scan was not<br />

affected by sighting conditions. There was no change in the mean number of manatees sighted<br />

per scan from 2001 to the 2007. Our ability to detect a trend ranged from 9% to 100%<br />

depending on the level of decline, scan duration, number of points surveyed and number of<br />

surveys. This survey protocol is a practical and repeatable way to examine population trends of<br />

sirenians in similar habitats around the world.<br />

Keywords: Monitoring, trend detection, power analysis, West Indian manatees, Trichechus sp.,<br />

sirenians, boat surveys.<br />

*Corresponding author: Tel. (248) 756-3985. E-mail addresses: kslacommare@gmail.com.<br />

Abbreviations: IUCN- <strong>International</strong> Union for the Conservation of Nature; SPAW- Specially Protected<br />

Areas and Wildlife; CITES- Convention on <strong>International</strong> Trade of Endangered Species, MMPA <strong>–</strong> Marine<br />

Mammal Protection Act<br />

1


1
<br />

2
<br />

3
<br />

4
<br />

5
<br />

6
<br />

7
<br />

8
<br />

9
<br />

10
<br />

11
<br />

12
<br />

13
<br />

14
<br />

15
<br />

16
<br />

<strong>17</strong>
<br />

18
<br />

19
<br />

20
<br />

21
<br />

22
<br />

1 Introduction<br />

All four extant species in the order Sirenia are vulnerable to extinction due to small<br />

population sizes, population declines, fragmentation and continued exploitation (Deutsch et<br />

al. 2010; Lefebvre et al. 2001; Marsh and Lefebvre 1994). They are listed as vulnerable on<br />

the IUCN Red List (Deutsch et al. 2010), protected by CITES (CITES 2010), the SPAW<br />

protocol (CEP 2010), the Memorandum of Understanding on the Conservation and<br />

Management of Dugongs (CMS 2010), the US Endangered Species Act (USFWS 2001),<br />

and other national and international wildlife protection laws (see Quintana-Rizzo and<br />

Reynolds 2008). For many sirenian populations, little information exists on their status<br />

(e.g. Hines et al. 2005; Lefebvre et al. 2001; Marsh and Lefebvre 1994; Quintana-Rizzo and<br />

Reynolds 2008). Monitoring <strong>–</strong> a preliminary step in determining conservation status,<br />

effects of exploitation and improving management decisions (Gibbs 2000; Gibbs et al.<br />

1998; Marsh and Trenham 2008; Martin et al. 2006) <strong>–</strong> is the process of gathering data to<br />

draw inferences about population abundance changes over time (Yoccoz et al. 2001).<br />

Monitoring is listed as a key objective in most manatee (Trichechus spp) and dugong<br />

(Dugong dugon) conservation and management plans (eg. Auil 1998; CEP 1995; USFWS<br />

2001).<br />

As others have pointed out, “the need for high quality data is clear, the means for<br />

getting it is not (Dawson et al. 2008, p. 20; USFWS 2001),” especially for small<br />

populations of marine mammals that live in complex coastal and riverine habitats <strong>–</strong><br />

conditions that constrain survey designs (see Dawson et al. 2004; Dawson et al. 2008; Dick<br />

and Hines 2011; Hines et al. 2005; Williams and Thomas 2009). A well designed<br />

2


23
<br />

24
<br />

25
<br />

26
<br />

27
<br />

28
<br />

29
<br />

30
<br />

31
<br />

32
<br />

33
<br />

34
<br />

35
<br />

36
<br />

37
<br />

38
<br />

39
<br />

40
<br />

41
<br />

42
<br />

43
<br />

44
<br />

45
<br />

monitoring program requires two key components: a reliable population index and<br />

powerful statistical analysis. Indices are a proxy for population abundance (Caughley<br />

1977; Gibbs 2000; Kindberga et al. 2009) and are based on the premise that systematic<br />

surveys will detect the same proportion of the population over time. Thus, changes in the<br />

number of animals detected reflect changes in population size (Gibbs 2000). To be a good<br />

surrogate to population size, an index must have a positive, linear relationship with actual<br />

abundance (Gibbs 2000; Gibbs et al. 1998; Williams and Thomas 2009), and a constant<br />

detection probability over habitat, sighting conditions and time (Anderson 2001; Gibbs<br />

2000; Gibbs et al. 1998; Thompson 2004; Williams and Thomas 2009).<br />

Detecting trends in abundance depends on statistical power (Gerrodette 1987; Gibbs<br />

2000; Gibbs et al. 1998; Taylor and Gerrodette 1993) and an effective monitoring program<br />

must generate data that can be statistically analyzed to detect trends (Gibbs 2000; Gibbs et<br />

al. 1998). Both the accuracy of the population index and sampling structure <strong>–</strong> number of<br />

plots, survey frequency, scan duration, number of years <strong>–</strong> will affect the ability to detect<br />

changes in population abundance. Power analysis, through simulation, is the process used<br />

to determine the statistical power of an index and sampling regime (Gerrodette 1987; Gibbs<br />

2000; Gibbs et al. 1998; Taylor and Gerrodette 1993).<br />

Most sirenian populations are found in developing nations where monitoring funds<br />

are scarce. To be valuable, monitoring methods must be sound, repeatable and inexpensive<br />

(Aragones et al. in press; Aragones et al. 1997; Dick and Hines 2011; Hines et al. 2005;<br />

Williams and Thomas 2009). This paper evaluates a low cost and repeatable boat-based<br />

method for monitoring sirenians. We used this method on West Indian manatees<br />

(Trichechus manatus manatus) in the Drowned Cayes area of Belize, Central America and<br />

3


46
<br />

47
<br />

48
<br />

49
<br />

50
<br />

51
<br />

52
<br />

53
<br />

54
<br />

55
<br />

56
<br />

57
<br />

58
<br />

59
<br />

60
<br />

61
<br />

62
<br />

63
<br />

64
<br />

65
<br />

66
<br />

examined trends in manatee abundance from 2001-2007. Through simulations, we<br />

conducted a power analysis to determine trend detection ability and the ramifications of<br />

sampling structure on trend detection.<br />

2. Materials and Methods<br />

2.1 Study area<br />

The Drowned Cayes, including Swallow Caye, are a string of mangrove islands -14 km<br />

long by 4 km wide - along the central coast of Belize, 10-15 km east of Belize City and 3-5<br />

km west of the Belize Barrier reef (Figure 1). These islands are surrounded by shallow<br />

seagrass beds. The depth of the study area is less than 1 meter in some places and is never<br />

greater than 6 meters. The study area also encompasses two points along the Belize Barrier<br />

Reef. Prior to this study, aerial and boat surveys and DNA analyses documented that this<br />

area has an important population of manatees (Auil 2004; Bengston and Magor 1979;<br />

Hunter et al. 2010; LaCommare et al. 2008, Morales-Vela et al. 2000; O'Shea and Salisbury<br />

1991, Self-Sullivan 2008).<br />

This study area is typical of manatee coastal habitat making it appropriate for<br />

testing our survey protocol. It comprises: shallow water (1-3m deep), proximity to deeper<br />

channels (3-6 m) for escape from danger (boats), freshwater sources, aquatic vegetation<br />

(mostly submerged), warm water (<strong>17</strong>°C minimum temperature), shelter from storms,<br />

waves, strong winds and currents; and travel corridors between habitat areas (Moraes-<br />

Arraut et al. in press).<br />

4


67
<br />

68
<br />

69
<br />

70
<br />

71
<br />

72
<br />

73
<br />

74
<br />

75
<br />

76
<br />

77
<br />

78
<br />

79
<br />

80
<br />

81
<br />

82
<br />

83
<br />

84
<br />

85
<br />

86
<br />

87
<br />

88
<br />

2.2 Survey design<br />

We devised a point-based survey design to monitor manatee occurrence by counting the<br />

number of manatees sighted during a 20-minute period at designated locations throughout<br />

the Drowned Cayes area (Figure 1). We marked 54 permanent points throughout the study<br />

area with a global positioning system (GPS) and used a small (8m) fiberglass boat to survey<br />

the points on a regular basis. To minimize boat and engine-noise disturbance to the<br />

manatees, we maneuvered the boat with a 8-meter pole when we were within 100 meters of<br />

the exact point location. During this time, 3 - 13 observers started scanning for manatees.<br />

At least three experienced observers were on the boat at all times. Experienced observers<br />

included one of the two of the authors (KSL and CSS), a field assistant that worked with<br />

the authors for the entire study and student interns that spent 3 <strong>–</strong> 6 months at the field site.<br />

Student interns were trained to spot manatees <strong>–</strong> nose, tail, back - and signs of manatees -<br />

shadows under the water, tail pressure wave, disturbed silt, floating grass -by the authors<br />

and field assistant. The number of inexperienced, but trained observers (volunteers),<br />

ranged from 0 to 10. Inexperienced observers were trained to spot manatees and manatee<br />

signs by the authors, field assistants and interns at the beginning of each two-week session.<br />

Volunteers were introduced to manatee viewing first via a PowerPoint lecture and then<br />

asked to practice spotting and viewing live manatees by visiting Swallow Caye Wildlife<br />

Sanctuary <strong>–</strong> a local manatee sanctuary and viewing area. Once anchored in position with<br />

the pole, observers searched for manatees in a 360 o circle around the boat. Having at least<br />

three experienced manatee spotters on the boat at all times ensured that experienced<br />

observers could cover the entire 360 degree circle and not rely on volunteers to detect<br />

5


89
<br />

90
<br />

91
<br />

92
<br />

93
<br />

94
<br />

95
<br />

96
<br />

97
<br />

98
<br />

99
<br />

100
<br />

101
<br />

102
<br />

103
<br />

104
<br />

105
<br />

106
<br />

107
<br />

108
<br />

109
<br />

110
<br />

manatees. A manatee was counted if a tail, back, nose or entire manatee was spotted. If a<br />

manatee was spotted by an inexperienced observer, it was only counted if it was confirmed<br />

by an experienced observer. From 2001-2007, we conducted these point scans 4 - 8 times<br />

per year during two-week survey sessions and sampled the points without replacement for<br />

the two-week period. Since not all points could be surveyed in one day, we divided the<br />

study area into 8 zones with 4 <strong>–</strong> 6 points each. Each day, we randomly chose the survey<br />

zone and starting point and then surveyed 4 <strong>–</strong> 5 points.<br />

2.3 Index and detection probability<br />

Our population index was the number of manatees sighted per 20-minutes per point. A<br />

valid population index must have a constant, linear relationship with manatee abundance.<br />

Two conditions can violate this assumption. When using call indices (indices that rely on<br />

animal calls such as frog calls or bird songs) or presence/absence data (Gibbs 2000; Gibbs<br />

et al. 1998), high population densities have an asymptotic relationship between the index<br />

and abundance. Alternatively, low population densities can have a threshold below which<br />

animals will not be detected (Gibbs 2000; Gibbs et al. 1998).<br />

We were not using presence/absence data or call indices, and since we were<br />

evaluating population declines rather than population increases, we were more concerned<br />

with the problem of a bottom threshold than population saturation. This point-based<br />

method has been successfully used to locate and count small and sparse populations of<br />

manatees in: Lake Volta, Ghana; Estero Hondo, Dominican Republic; and Lake Ossa,<br />

Cameroon (Self-Sullivan, personal communication; Dominquez, personal communication).<br />

This method can detect manatees at low population densities. To further evaluate this<br />

6


111
<br />

112
<br />

113
<br />

114
<br />

115
<br />

116
<br />

1<strong>17</strong>
<br />

118
<br />

119
<br />

120
<br />

121
<br />

122
<br />

123
<br />

124
<br />

125
<br />

126
<br />

127
<br />

128
<br />

129
<br />

130
<br />

131
<br />

132
<br />

133
<br />

assumption, we divided our locations into high-sighting probability sites (>30% probability<br />

of sighting a manatee per scan) and low-sighting probability sites (< 30% probability of<br />

sighting a manatee per scan) (LaCommare et al. 2008). We then examined cumulative<br />

histograms of time to first sighting for both categories. We expected that each histogram<br />

would have a similar asymptotic curve indicating that our ability to detect manatees was<br />

similar for high versus low-sighting probability locations.<br />

The relationship of the index to the actual number of animals counted is also a<br />

function of detection probability (Anderson 2001; Gibbs 2000; Gibbs et al. 1998;<br />

Thompson 2004; Williams and Thomas 2009) <strong>–</strong> the ability to detect an animal when it is<br />

present (MacKenzie and Royle 2005). Three classes of variables affect this: observers,<br />

environment and species behavior. For each point scan, we counted the number of<br />

manatees and recorded variables that might influence detection probabilities. We used a<br />

generalized linear model (GLM) with a negative binomial distribution and a log-link<br />

function to determine whether the number of manatees sighted was influenced by sighting<br />

conditions. This distribution was the most appropriate for our over-dispersed (variance<br />

greater than the mean) Poisson-distributed response variable, number of manatees (Agresti<br />

1996; Quinn and Keough 2002). Our predictor variables were: environmental conditions<br />

(sun glare - yes/no, precipitation -dry/light rain, cloud cover -clear/scattered clouds/partly<br />

cloudy/overcast, Beaufort sea state -0/1/2/3, swell height -in 0.15m increments; water<br />

clarity and time of day), and manatee behavior (disturbed/feeding/resting/social/travel/<br />

undetermined - Table 1). Water clarity was a measure of the horizontal distance between<br />

an underwater observer and Secchi disk held 0.5 meters under the water. Water clarity was<br />

measured in this fashion because the vertical Secchi depth was usually greater than or equal<br />

7


134
<br />

135
<br />

136
<br />

137
<br />

138
<br />

139
<br />

140
<br />

141
<br />

142
<br />

143
<br />

144
<br />

145
<br />

146
<br />

147
<br />

148
<br />

149
<br />

150
<br />

151
<br />

152
<br />

153
<br />

154
<br />

to the depth. We did not take seasonality into account because previous analyses indicated<br />

that there is no difference in seasonal sighting probability within the Cayes (LaCommare et<br />

al. 2005; Self-Sullivan 2008).<br />

To test whether detection probability was influenced by number and experience of<br />

volunteers we used a generalized linear model (GLM) with a negative binomial distribution<br />

and a log-link function to determine whether the number of manatees sighted was<br />

influenced by number of volunteers and year (Table 2).<br />

2.4 Trends of manatee counts in the Drowned Cayes<br />

Forty-two survey periods were conducted over 79 months from 2001-2007. We used a<br />

GLM with negative binomial distribution and a log-link function to determine whether<br />

there was a change in the number of manatees per 20-minute scan from the first survey<br />

period in 2001 to the last survey period in 2007.<br />

2.5 Power analysis of trend data<br />

We used simulations, following Taylor and Gerrodette (1993) and Gibbs (1998, 2000), to<br />

determine the statistical power of our ability to detect a trend. We fit a negative binomial<br />

distribution to our manatee count data from our 20-minute scans for the whole 2001 - 2007<br />

period. Using the parameters from this distribution, we generated a random array of the<br />

number of manatees for 28 points for each of 6 survey periods per year. The potential<br />

number of manatees sighted was truncated at 5 since we only had 1 sighting with more than<br />

5 manatees in 7 years. We chose this sampling regime because this most closely matched<br />

our annual survey structure from 2001 <strong>–</strong> 2007. We repeated this process for 15 “years”.<br />

8


155
<br />

156
<br />

157
<br />

158
<br />

159
<br />

160
<br />

161
<br />

162
<br />

163
<br />

164
<br />

165
<br />

166
<br />

167
<br />

168
<br />

169
<br />

<strong>17</strong>0
<br />

<strong>17</strong>1
<br />

<strong>17</strong>2
<br />

<strong>17</strong>3
<br />

<strong>17</strong>4
<br />

<strong>17</strong>5
<br />

Then we examined whether we could detect a decline in the number of manatees over that<br />

15 year period. We generated our random array, added a linear trend, and ran our GLM<br />

1000 times to determine our ability to detect each of three declines in the number of<br />

manatees at the end of these 15 years: slight (10%), moderate (25%) and precipitous (50%).<br />

These declines are equivalent to a 0.7%, 1.7% and 3.6% annual decrease in the number of<br />

manatees sighted per 20-minute scan. We considered that a negative trend was detected if<br />

the slope of the model in each simulation run was negative and significantly different from<br />

zero. Following Taylor et al. (2007), we chose a 15 year period for our power analysis.<br />

2.6 Power analysis of sampling design<br />

Survey structure affects variability in detection probability and therefore statistical power<br />

(Gibbs 1998, 2000). Number of plots, sampling frequency, sampling interval and scan<br />

duration interact to influence statistical power and need to be considered when designing a<br />

monitoring protocol. These factors also influence the cost and time of monitoring.<br />

Therefore, determining the most efficient experimental design can be critical to carrying<br />

out an effective monitoring program. In the field, we sampled approximately 28 of our 54<br />

points during two-week survey periods, 4-8 times per year. Each point was scanned for a<br />

minimum of 20 minutes. A fraction of the point scans lasted 30 minutes, but this decreased<br />

the number of points that could be sampled in any one day and during a two-week survey<br />

period. How does a 30-minute scan duration with fewer samples impact our ability to<br />

detect a trend? Using data that we collected from our 30-minute scans, we generated new<br />

negative binomial distribution parameters and a new random array of the number of<br />

9


<strong>17</strong>6
<br />

<strong>17</strong>7
<br />

<strong>17</strong>8
<br />

<strong>17</strong>9
<br />

180
<br />

181
<br />

182
<br />

183
<br />

184
<br />

185
<br />

186
<br />

187
<br />

188
<br />

189
<br />

190
<br />

191
<br />

192
<br />

193
<br />

194
<br />

195
<br />

196
<br />

manatees for 21 sampling points per two-week survey period for 6 survey periods per year<br />

and repeated our power analysis as described above.<br />

In the Drowned Cayes, 28 out of 54 point scan locations have a greater than 30%<br />

probability of sighting a manatee (LaCommare et. al, 2008). Would restricting our survey<br />

protocol to locations with a higher sighting probability yield a greater ability to detect a<br />

trend? We generated new negative binomial distribution parameters using our manatee<br />

count data from 2001 - 2007 from locations with a greater than 30% probability of sighting<br />

a manatee (LaCommare et. al 2008). Using these parameters, we generated a new random<br />

array of the number of manatees for 28 sampling points for 6 two-week survey periods.<br />

We repeated our power analysis as described above. From that distribution, we also created<br />

an array of the number of manatees for 8 two-week survey periods with both 28 and 21<br />

points and for 8 two-week survey periods, 28 points with just a 20-minute scan duration.<br />

3. Results<br />

3.1 Index and encounter probability<br />

Histograms of time-to-first sighting indicate that for locations that had a greater than 30%<br />

probability of sighting a manatee, 90% of all manatees are sighted within the first 20<br />

minutes of a 30-minute scan (Figure 2). For locations that had a less than 30% probability<br />

of sighting a manatee, it took slightly longer to spot the first manatee and 90% of all<br />

manatees were sighted within 24 rather than 20 minutes (Figure 2).<br />

The number of manatees sighted per scan was not affected by sighting conditions<br />

(overall Likelihood Ratio Chi-Square = 12.578, df = 18, p = 0.816, n= 131, Table 1). None<br />

10


197
<br />

198
<br />

199
<br />

200
<br />

201
<br />

202
<br />

203
<br />

204
<br />

205
<br />

206
<br />

207
<br />

208
<br />

209
<br />

210
<br />

211
<br />

212
<br />

213
<br />

214
<br />

215
<br />

216
<br />

2<strong>17</strong>
<br />

of the individual factors or covariates had a significant influence on the number of<br />

manatees sighted per scan (Table 1).<br />

There was a negative effect between number of volunteers and number of manatees<br />

spotted (Wald Chi-Square = 7.343 df =1, p < 0.007, Table 2). The greater number of<br />

volunteers present on the boat, the fewer number of manatees spotted.<br />

3.2 Trends of manatee counts in the Drowned Cayes<br />

We conducted 960 20-minute scans during 42 survey periods from 2001 - 2007. The<br />

number of manatees sighted per scan ranged from 0 to 6. Overall mean number of<br />

manatees sighted per scan was 0.54 (0.89 SD) (Figure 3). Mean number of manatees per<br />

survey period ranged from 0.21 to 1.08 manatees per scan. Yearly means per scans ranged<br />

from 0.33 manatees in 2002 to 0.71 manatees in 2001 (Figure 4). There was no detectable<br />

change in the mean number of manatees sighted per scan from the first survey period in<br />

2001 to the last survey period in 2007 (Likelihood Ratio Chi-Square = 1.56, df = 1, p =<br />

0.212, B = -0.003, Figure 4).<br />

3.3 Power Analysis<br />

Using a 20-minute scan duration, 28 points per survey period, and 6 periods, we had 4%<br />

power to detect a 10% decline in number of manatees over 15 years within 7 years <strong>–</strong> our<br />

study duration, 9% power to detect a 25% decline and 22% power to detect a 50% decline.<br />

Extending this study duration to 15 years would give us a 11% power to detect a 10%<br />

decline, 51% power to detect a 25% decline and 100% power to detect a 50% decline<br />

(Table 3; Figure 5a, b, c).<br />

11


218
<br />

219
<br />

220
<br />

221
<br />

222
<br />

223
<br />

224
<br />

225
<br />

226
<br />

227
<br />

228
<br />

229
<br />

230
<br />

231
<br />

232
<br />

233
<br />

234
<br />

235
<br />

236
<br />

237
<br />

238
<br />

3.4 Power Analysis of sampling design<br />

The highest power, i.e. 16%, 71% and 100% after 15 years, is achieved for all three decline<br />

rates when 28 points with a high (>30%) probability of sighting a manatee are surveyed 8<br />

times per year (Table 3). This survey design has a 90% power to detect a 50% decline by<br />

the 11 th year (Figure 5c). The design with the least power, i.e. 8%, 34% and 99% after 15<br />

years, surveys just 21 points, regardless of the probability of sighting a manatee, 6 times a<br />

year (Table 3). Overall, survey protocol changes had little impact on our ability to detect<br />

slight declines over 15 years, but did impact our ability to detect moderate and precipitous<br />

declines (Figure 5a, b, c) over that time horizon. Increasing the number of scan points from<br />

21 to 28 had about the same effect as increasing the scan duration from 20 to 30 minutes<br />

(about a 10% increase in power to detect moderate declines). Increasing the number of<br />

survey periods from 6 to 8 resulted in an increase in power from 57% to 71% in the<br />

moderate decline scenario (Table 3). Differences in power across designs are not observed<br />

before 10 years of survey in the case of slight declines, 8 years for moderate declines, and 6<br />

years for precipitous declines (Figure 5a, b, c). All designs show the same basic result:<br />

only precipitous declines can be detected with a 100% certainty on a 15 year time horizon<br />

(Figure 5a, b, c).<br />

4. Discussion<br />

Using 30-minute point scans from a small boat platform is an effective and repeatable<br />

method for monitoring manatees in typical manatee habitats around the world. Several<br />

authors debate the validity of population indices in comparison to population estimates - the<br />

12


239
<br />

240
<br />

241
<br />

242
<br />

243
<br />

244
<br />

245
<br />

246
<br />

247
<br />

248
<br />

249
<br />

250
<br />

251
<br />

252
<br />

253
<br />

254
<br />

255
<br />

256
<br />

257
<br />

258
<br />

259
<br />

260
<br />

261
<br />

former characterized as estimating relative abundance, the latter absolute abundance (see<br />

Thompson 2004; Williams and Thomas 2009). These authors caution against the use of<br />

indices in favor of more robust estimating procedures such as distance sampling or mark-<br />

recapture procedures (e.g. Anderson 2001; Williams and Thomas 2009). Yet, as long as the<br />

population index is linearly related to actual abundance <strong>–</strong> that is it doesn’t have an<br />

asymptote or bottom threshold, and detection probabilities are constant over space and<br />

time, count-based indices are valid (Skalski et al. 1983; Thompson 2004; Williams and<br />

Thomas 2009).<br />

Species behavior <strong>–</strong> long dive times, short surfacing bouts - and severely limited<br />

resources hamper the ability to conduct robust monitoring procedures on Antillean<br />

manatees and other sirenians in developing nations (Aragones et al. in press; Aragones et<br />

al. 1997; Dawson et al. 2008; Dick and Hines 2011). Distance sampling requires the ability<br />

to consistently measure the distance from the observer to the animal using a range finder,<br />

reticulated binoculars or spotting scope (Buckland et al. 2001). <strong>Sirenian</strong>s often surface<br />

very briefly and only expose their nose to the surface when breathing. In our study, even<br />

the best trained observers captured distances via range finder only 10% of the time<br />

(LaCommare and Self-Sullivan, personal observation). Mark-recapture methods are<br />

equally challenging. It is estimated that 50% of manatees in the study area are scarred by<br />

boats (Self-Sullivan 2008). Although possible (Self-Sullivan 2008; Auil and Powell,<br />

personal communication), above-water photo identification methods are difficult because<br />

manatees in Belize and other tropical habitats do not seasonally aggregate at warm water<br />

sites, nor do they tend to float at the surface, exposing their scarred backs. Therefore,<br />

photo-identification via underwater video capturing techniques is more appropriate.<br />

13


262
<br />

263
<br />

264
<br />

265
<br />

266
<br />

267
<br />

268
<br />

269
<br />

270
<br />

271
<br />

272
<br />

273
<br />

274
<br />

275
<br />

276
<br />

277
<br />

278
<br />

279
<br />

280
<br />

281
<br />

282
<br />

283
<br />

284
<br />

However, this involves considerable time and effort by the observer and does not always<br />

yield individual identification due to lack of natural markings and/or poor visibility in many<br />

habitats (LaCommare and Self-Sullivan, personal observation). Preliminary analysis<br />

indicates that population abundance estimates via mark-recapture techniques might be<br />

possible (Self-Sullivan 2008, Auil & Powell, Personal communication). Distance sampling<br />

and photo-identification requires considerable observer training. Equipment and observer<br />

training may price monitoring programs out of reach of most wildlife and conservation<br />

agencies in developing nations (Aragones et al. in press; Aragones et al. 1997).<br />

Our index does not appear to have a bottom threshold below which manatees are not<br />

detected, but 30-minutes is a better scan duration than 20-minutes. The histograms of time-<br />

to-first-sighting indicate that for locations with a greater than 30% probability of sighting a<br />

manatee, roughly 90% of all manatees are sighted within the first 20 minutes of a 30-<br />

minute scan. But, for locations with a less than 30% probability of sighting a manatee, it<br />

takes longer to spot the first manatee and it takes up 24 minutes before 90% of all manatees<br />

are sighted.<br />

Based on our analysis of possible influences on detection probability, the<br />

relationship between our index and population abundance is constant over conditions and<br />

time because environmental conditions and manatee behavior do not appear to significantly<br />

influence the number of manatees sighted. This is a surprising result. Visibility bias, bias<br />

from the result of observers missing animals due to observer inconsistencies and<br />

environmental conditions, is significant in aerial surveys for manatees (See Packard et al.<br />

1985 and Wright et al. 2002). Sighting conditions such as water clarity, turbidity, sea state<br />

and surface glare have all been shown to affect visibility (i.e. Reynolds and Wilcox 1994)<br />

14


285
<br />

286
<br />

287
<br />

288
<br />

289
<br />

290
<br />

291
<br />

292
<br />

293
<br />

294
<br />

295
<br />

296
<br />

297
<br />

298
<br />

299
<br />

300
<br />

301
<br />

302
<br />

303
<br />

304
<br />

305
<br />

306
<br />

307
<br />

in aerial surveys. There are two principal reasons why these factors may not affect<br />

visibility of manatees from boats. From a boat, observers search for manatees across the<br />

horizontal plane of the water looking for signs of the manatees’ body (a nose, tail or back)<br />

breaking through the surface of the water. Water clarity isn’t going to affect this to the<br />

same extent as aerial surveys in which observers are searching for manatees within the<br />

water column. In extremely turbid water, like that found in “black water” lake and river<br />

systems of the Amazon, lower vantage points allow better visibility of an animal breaking<br />

the surface of the water than do higher vantage points (Aragones et al. in press). Sea state<br />

may not impact visibility because when surface choppiness is slight <strong>–</strong> sea state of zero or<br />

one <strong>–</strong> any disturbance to the surface of the water is noticeable and therefore manatees will<br />

be spotted. Because manatees need to rise their nose above the water to breathe, when<br />

there is a sea state of two or three <strong>–</strong> choppy conditions <strong>–</strong> manatees will raise their entire<br />

head out of the water (Personal observation, KSL and CSS) which facilitates the observer’s<br />

ability to spot manatees and negates the impact of sea state on visibility. Surveys were not<br />

conducted in a sea state higher than three. It is not clear why surface glare does not impact<br />

visibility. There are three possible explanations. Surveys are conducted between 9:30 am<br />

and 4:30 pm when surface glare, due to a sun low on the horizon, is at a minimum.<br />

Multiple trained observers searching for manatees may reduce the negative effects of sun<br />

glare. And, manatees usually surface multiple times during a 30-minute search duration<br />

which increases the possibility of spotting a manatee even during difficult observation<br />

conditions.<br />

The biggest concern with our study was that number of volunteers had a negative<br />

effect on our ability to spot manatees. The greater the number of volunteers the fewer<br />

15


308
<br />

309
<br />

310
<br />

311
<br />

312
<br />

313
<br />

314
<br />

315
<br />

316
<br />

3<strong>17</strong>
<br />

318
<br />

319
<br />

320
<br />

321
<br />

322
<br />

323
<br />

324
<br />

325
<br />

326
<br />

327
<br />

328
<br />

329
<br />

330
<br />

number of manatees spotted. Volunteers are easily distracted by each other during long<br />

scan periods and survey days and do not have the same level of focus as trained observers.<br />

This can easily be minimized in the future by limiting the number of volunteers present<br />

during surveys.<br />

We do not know the relationship of our index to actual abundance. It is an index of<br />

relative abundance that can be used to detect trends. Index validation, by comparing the<br />

number of manatees counted via boat surveys to the number counted during concurrent,<br />

localized aerial surveys, could provide an indication of the relationship between boat counts<br />

and aerial survey counts which could help determine bias and thus the relationship of our<br />

index to actual abundance. Aerial surveys were conducted along the entire Belizean<br />

coastline, including the Drowned Cayes, from 1997 <strong>–</strong> 2002 (Auil, 2004). These surveys<br />

overlapped with our boat surveys in 2000 - 2002, but were only conducted once during the<br />

wet and once during the dry season and did not directly correspond to our survey effort and<br />

unfortunately, are not useful for evaluating bias of our method. It is interesting to note that<br />

the trends of Auil’s (2004) index of relative abundance from 2000-2002 mirrors our trends<br />

across those same years.<br />

There was a very slight negative trend in the data, but this trend was not significant.<br />

Our analysis indicates that we would have only had a 4%, 9% and 22% power to detect<br />

slight, moderate or precipitous declines after 7 years <strong>–</strong> our study duration. Even slight<br />

negative trends can be important for small populations which can be impacted by stochastic<br />

events or behavioral biology that can drive small populations to extinction (Meffe and<br />

Carroll 1997) or prevent these populations from recovering (Meffe and Carroll 1997). The<br />

Antillean subspecies of the West Indian Manatee is currently listed as endangered on the<br />

16


331
<br />

332
<br />

333
<br />

334
<br />

335
<br />

336
<br />

337
<br />

338
<br />

339
<br />

340
<br />

341
<br />

342
<br />

343
<br />

344
<br />

345
<br />

346
<br />

347
<br />

348
<br />

349
<br />

350
<br />

351
<br />

352
<br />

353
<br />

IUCN Red List because it is a small and declining population without ongoing, effective<br />

conservation actions (Self Sullivan et al. 2011). These data provide confirmation of the<br />

appropriateness of the current IUCN status and highlight the need to improve our ability to<br />

establish the magnitude of decline.<br />

Extending our study to 15 years considerably improves our ability to detect a<br />

population trend. Over a study duration of 15 years, we would still have a limited ability to<br />

detect slight and moderate declines, but our index and protocol can detect precipitous<br />

declines within the benchmarks established in the Marine Mammal Protection Act, the<br />

IUCN Red List (Taylor et al 2007) and other publications. Taylor et al. (2007) defined a<br />

precipitous decline in marine mammal abundance as a 50 percent decline after 15 years<br />

because a 50% decline over this period would result in a stock being classified as<br />

“depleted” under the MMPA. Under the IUCN Red List guidelines, a decline of this<br />

magnitude could result in a status designation anywhere between “vulnerable” to<br />

“endangered” depending on initial population size, cause, certainty and the reversibility of<br />

the decline (IUCN 2010). IUCN Red List guidelines establish level of endangerment <strong>–</strong><br />

nearly threatened to critically endangered <strong>–</strong> based on declines of 10 <strong>–</strong> 90% over 10 years or<br />

3 generations (whichever is longer), depending on initial population size, cause, certainty<br />

and the reversibility of the decline (IUCN 2010). Bart et al. (2004) recommended a<br />

standard for landbird surveys of an 80% power to detect a 50% decline over 20 years.<br />

Similarly, Hatch (2003) recommended a standard for counts of colonial seabirds of a 90%<br />

power to detect a 50% decline over 10 years. Using this sampling regime over a 15 year<br />

time interval, we would have a power of 11%, 51% and 100% to detect slight, moderate or<br />

precipitous declines. All of our survey protocols had 100% power to detect 50% declines<br />

<strong>17</strong>


354
<br />

355
<br />

356
<br />

357
<br />

358
<br />

359
<br />

360
<br />

361
<br />

362
<br />

363
<br />

364
<br />

365
<br />

366
<br />

367
<br />

368
<br />

369
<br />

370
<br />

371
<br />

372
<br />

373
<br />

374
<br />

375
<br />

over 15 years making it a valuable tool for determining population status based on the<br />

benchmarks discussed above. Taylor et al. (2007) found that most marine mammal<br />

monitoring programs had a low power (0-50%) to detect precipitous declines in abundance<br />

within this time frame.<br />

We can improve the power of our survey protocol by surveying at least 28 points, 8<br />

times per year, increasing our scan duration to 30-minutes and choosing locations with a<br />

30% sighting probability. Under this regime, we have 100% power to detect a 50% decline<br />

over a 15 year period. We have a power of 90% to detect a 50% decline after 10 years <strong>–</strong><br />

exceeding the MMPA benchmark. The power of this sampling regime also meets the<br />

IUCN guidelines for establishing whether a population should be listed as critically<br />

endangered, endangered or vulnerable (IUCN, 2010). We would still have a limited ability<br />

to detect slight (16% power) declines, but a reasonable ability (71% power) to detect<br />

moderate declines. The IUCN benchmarks for establishing the vulnerability status of small<br />

populations for whom the causes and reversibility of declines are uncertain range from 10 <strong>–</strong><br />

30 % declines over 3 generations (60 years for manatees, (Deutsch et al. 2008)).<br />

To ensure the greatest success in a monitoring program, pilot studies are highly<br />

recommended (Aragones et al. in press; Buckland et al. 2001; Gibbs 2000; Gibbs et al.<br />

1998; Scheiner and Gurevitch 2001; Thompson 2004). Local knowledge can be utilized in<br />

pilot studies to locate high use areas in a short period of time <strong>–</strong> e.g. 1 year (Self-Sullivan<br />

and LaCommare, personal observation). Including local knowledge in research programs<br />

also has the effect of improving success in corollary conservation programs - e.g. poaching<br />

reduction programs (Aragones et al. in press; Kendall 2009). In addition to being a<br />

18


376
<br />

377
<br />

378
<br />

379
<br />

380
<br />

381
<br />

382
<br />

383
<br />

384
<br />

385
<br />

386
<br />

387
<br />

388
<br />

389
<br />

390
<br />

391
<br />

392
<br />

393
<br />

394
<br />

395
<br />

396
<br />

397
<br />

398
<br />

399
<br />

400
<br />

401
<br />

practical and repeatable way to examine population trends, boat surveys can be used to<br />

assess spatial distribution and habitat use (LaCommare et al. 2008).<br />

4. Conclusion<br />

Using a small-boat platform for systematic surveys is a sound and repeatable<br />

method for monitoring sirenian populations in developing nations where constraints on<br />

fiscal resources are severe. We recommend pilot studies and using local knowledge to<br />

locate high use areas and then using a 30-minute scan duration to conduct point scans. Our<br />

analysis indicates that surveys should be conducted at least 8 times per year and at least 21<br />

points should be included in the sampling protocol. For our study area, we estimate this<br />

monitoring protocol would require approximately 37 dedicated days on the water per year<br />

(4-5 scans per day, 3 days per month) with the only expenses being boat, fuel, and three<br />

well-trained observers. Alternatively, this protocol would be well suited for the eco-<br />

tourism industry where well trained local tour guide/researchers could conduct one or two<br />

scans during each tour, which are randomly distributed among 21 sites with a >30%<br />

sighting probability.<br />

Acknowledgements. We thank Earthwatch Institute for substantial financial support and<br />

Earthwatch Volunteers for logistical support and data collection. We thank Conservation<br />

Action Fund for financial support that specifically supported capacity building among local<br />

field assistants and tour guides. We are ever grateful to Mr. Sidney Turton and Spanish<br />

Bay Resort, who provided logistical and in kind support from 2001-2004, and to Ms.<br />

Teresa Parkey and The Hugh Parkey Foundation for Marine Awareness and Education,<br />

who provided logistical and in kind support in 2005 - 2007. Supplemental financial support<br />

was provided by two individual Earthwatch volunteers (Mr. Greenwalt, amd Mr Burtt),<br />

Project Aware, Virtual Explorers and the Lerner-Gray Marine Research Fund (AMNH).<br />

We are grateful to the numerous interns and field assistants who were vital to data<br />

collection and field logistics. One author (CSS) was also supported by an NSF Graduate<br />

19


402
<br />

403
<br />

404
<br />

405
<br />

406
<br />

Fellowship from 2001-2003. We are especially thankful for the help and guidance of our<br />

colleague, friend, primary field assistant and source of traditional knowledge, Mr. Gilroy<br />

Robinson. Finally, we thank two anonymous reviewers whose comments greatly improved<br />

this paper.<br />

20


406
<br />

407
<br />

408
<br />

409
<br />

410
<br />

411
<br />

412
<br />

413
<br />

414
<br />

415
<br />

416
<br />

4<strong>17</strong>
<br />

418
<br />

419
<br />

420
<br />

421
<br />

422
<br />

423
<br />

424
<br />

425
<br />

426
<br />

427
<br />

428
<br />

429
<br />

430
<br />

431
<br />

432
<br />

433
<br />

434
<br />

435
<br />

6. Literature Cited<br />

Agresti, A., 1996. An Introduction to Categorical Data Analysis. Wiley, New York.<br />

Anderson, D.R., 2001. The need to get the basics right in wildlife field studies. Wildlife<br />

Society Bulletin 29, 1294-1297.<br />

Aragones, L., LaCommare, K.S., Kendall, S., Castelblanco-Martinez, D.N., Gonzalez-<br />

Socoloske, D., in press. Boat and land-based surveys for sirenians, In Hines, E., Reynolds,<br />

J., Mignucci-Giannoni, A, Aragones, L., and M. Marmontel (eds). <strong>Sirenian</strong> Conservation:<br />

Issues and Strategies in Developing Countries. University Press of Florida.<br />

Aragones, L.V., Jefferson, T.A., Marsh, H., 1997. Marine mammal survey techniques in<br />

developing countries. Asian Marine Biology 14, 15-39.<br />

Auil, N., 1998. Belize Manatee Recovery Plan, In UNDP/GEF Coastal Zone Management<br />

Project No. BZE/92/G31. United Nations Environment Program, Caribbean Environment<br />

Program, Kingston, Jamaica.<br />

Auil, N.E., 2004. Abundance and Distribution Trends of the West Indian Manatee in the<br />

Coastal Zone of Belize: Implication for Conservation, In Department of Fisheries and<br />

Wildlife. p. 83. Texas A&M University, College Station, TX.<br />

Bart, J., Burnham, K.P., Dunn, E.H., Francis, C.M., Ralph, C.J., 2004. Goals and strategies<br />

for estimating trends in landbird abundance. The Journal of Wildlife Management 68, 611-<br />

626.<br />

Bengston, J.L., Magor, D., 1979. A Survey of manatees in Belize. Journal of Mammalogy<br />

60, 230-232.<br />

Buckland, S.T., Anderson, D.R., Burnham, K.P., Laake, J.L., Borchers, D.L., Thomas, L.,<br />

2001. Introduction to Distance Sampling: Estimating abundance of biological populations.<br />

Oxford University Press, New York.<br />

Caughley, G., 1977. Analysis of Vertebrate Populations. Wiley-Interscience Publication.<br />

CEP, 1995. Regional Management Plan for the West Indian Manatee (Trichechus manatus<br />

manatus), In Technical Report No. 35. CEP UNEP, Kingston, Jamaica.<br />

CEP, 2010. Protocol Concerning Specially Protected Areas and Wildlife to the Convention<br />

for the Protection and Development of the Marine Environment of the Wider Caribbean<br />

21


436
<br />

437
<br />

438
<br />

439
<br />

440
<br />

441
<br />

442
<br />

443
<br />

444
<br />

445
<br />

446
<br />

447
<br />

448
<br />

449
<br />

450
<br />

451
<br />

452
<br />

453
<br />

454
<br />

455
<br />

456
<br />

457
<br />

458
<br />

459
<br />

460
<br />

461
<br />

462
<br />

463
<br />

Region. http://www.cep.unep.org/cartagena-convention/spaw-protocol/spaw-protocolen.pdf.<br />

Last Accessed: April 15, 2010<br />

CITES, 2010. Convention on <strong>International</strong> Trade in Endangered Species of Wild Fauna and<br />

Flora. http://www.cites.org/eng/disc/species.shtml. Last Accessed: April 15, 2010<br />

CMS, 2010. The Memorandum of Understanding on the conservation and management of<br />

dugongs (Dugong dugon) and their habitats throughout their range, ed. C.o.M. Species.<br />

http://www.cms.int/species/dugong/dugong_mou.htm. Last Accessed: April 15, 2010<br />

Dawson, S., Slooten, E., DuFresne, S., Wade, P., Clement, D., 2004. Small-boat surveys for<br />

coastal dolphins: line-transect surveys for Hector's dolphins (Cephalorhynchus hectori).<br />

Fishery Bulletin 102, 441-451.<br />

Dawson, S., Wade, P., Slooten, E., Barlow, J., 2008. Design and field methods for sighting<br />

surveys of cetaceans in coastal and riverine habitats. Mammal Review 38, 19-49.<br />

Deutsch, C.J., Self-Sullivan, C., Mignucci-Giannoni, A., 2008. Trichechus manatus<br />

manatus, In IUCN Red List of Threatened Species. IUCN 2010 Version 2010.1.<br />

. Last Accessed: 15 April 2010.<br />

Dick,
D.M.
and
E.M.
Hines.
2011.
Development
and
implementation
of
distance
sampling
<br />

techniques
to
determine
bottlenose
dolphin
(Tursiops
truncatus)
abundance
at
Turneffe
<br />

Atoll,
Belize.
Marine
Mammal
Science:
27(3):
606‐621.
<br />

Gerrodette, T., 1987. A power analysis for detecting trends. Ecology 68, 1364-1372.<br />

Gibbs, J.P., 2000. Monitoring Populations, In: L. Boitani and T. Fuller (eds) Research<br />

Techniques in Animal Ecology: Controversies and Consequences. pp. 212-252. Columbia<br />

University Press, New York.<br />

Gibbs, J.P., Droege, S., Eagle, P., 1998. Monitoring populations of plants and animals.<br />

BioScience 48, 935-940.<br />

Hatch, S.A., 2003. Statistical power for detecting trends with applications to seabird<br />

monitoring. Biological Conservation 111, 3<strong>17</strong>-329.<br />

Hines, E.M., Adulyanuskol, K., Duffus, D.A., 2005. Dugong abundance along the Adaman<br />

coast of Thailand. Marine Mammal Science 21, 536-549.<br />

22


464
<br />

465
<br />

466
<br />

467
<br />

468
<br />

469
<br />

470
<br />

471
<br />

472
<br />

473
<br />

474
<br />

475
<br />

476
<br />

477
<br />

478
<br />

479
<br />

480
<br />

481
<br />

482
<br />

483
<br />

484
<br />

485
<br />

486
<br />

487
<br />

488
<br />

489
<br />

490
<br />

491
<br />

492
<br />

493
<br />

Hunter, M.E., Auil-Gomez, N.E., Tucker, K.P., Bonde, R.K., Powell, J., McGuire, P.M.,<br />

2010. Low genetic variation and evidence of limited dispersal in the regionally important<br />

Belize manatee. Animal Conservation 13 592<strong>–</strong>602.<br />

Kendall, S., 2009. Making Science Matter, In <strong>International</strong> Marine Conservation Congress<br />

Annual Convention. Washington, DC.<br />

Kindberga, J., Ericssona, G., Swensonb, J.E., 2009. Monitoring rare or elusive large<br />

mammals using effort-corrected voluntary observers. Biological Conservation 142, 159-<br />

165.<br />

LaCommare, K.S., Self-Sullivan, C., Brault, S., 2005. Distribution, Seasonal Occurrence<br />

and Habitat Use of Antillean Manatees in the Drowned Cayes Area of Belize, Central<br />

America, In 16th Biennial Conference on the Biology of Marine Mammals. San Diego, CA.<br />

LaCommare, K.S., Self-Sullivan, C., Brault, S., 2008. Distribution and Habitat Use of<br />

Antillean Manatees (Trichechus manatus manatus) in the Drowned Cays Area of Belize,<br />

Central America. Aquatic Mammals 34, 35-43.<br />

Lefebvre, L.W., Marmontel, M., Reid, J.P., Rathbun, G.B., Domning, D.P., 2001. Status<br />

and biogeography of the West Indian manatee, In: C.A. Woods and F.E. Sergile (eds)<br />

Biogeography of the West Indies: Patterns and Perspectives. pp. 425-474. CRC Press,<br />

Boca Raton.<br />

MacKenzie, D.I., Royle, J.A., 2005. Designing occupancy studies: general advice and<br />

allocating survey effort. Journal of Applied Ecology 42, 1105-1114.<br />

Marsh, D.M., Trenham, P.C., 2008. Current trends in plant and animal population<br />

monitoring. Conservation Biology 22, 647.<br />

Marsh, H., Lefebvre, L.W., 1994. <strong>Sirenian</strong> status and conservation efforts. Aquatic<br />

Mammals 20, 155-155.<br />

Martin, J., Kitchens, W., Hines, J., 2006. Importance of well-designed monitoring programs<br />

for the conservation of endangered species: case study of the snail kite conservation<br />

Biology 21, 472-481.<br />

Meffe, G. K, Carroll, C. R. 1997. Principles of Conservation Biology. 2 nd Ed. Sinauer and<br />

Associates, Sunderland, MA.<br />

23


494
<br />

495
<br />

496
<br />

497
<br />

498
<br />

499
<br />

500
<br />

501
<br />

502
<br />

503
<br />

504
<br />

505
<br />

506
<br />

507
<br />

508
<br />

509
<br />

510
<br />

511
<br />

512
<br />

513
<br />

514
<br />

515
<br />

516
<br />

5<strong>17</strong>
<br />

518
<br />

519
<br />

520
<br />

521
<br />

522
<br />

523
<br />

Moraes-Arraut, E., Ortega-Argueta, A., Olivera-Gómez, D. L. in press. Delineating and<br />

Assessing Habitats for <strong>Sirenian</strong>s In Hines, E., Reynolds, J., Mignucci-Giannoni, A,<br />

Aragones, L., and M. Marmontel (eds). <strong>Sirenian</strong> Conservation: Issues and Strategies in<br />

Developing Countries. University Press of Florida.<br />

Morales-Vela, B., Olivera-Gomez, D., III, J.E.R., Rathbun, G.B., 2000. Distribution and<br />

habitat use by manatees (Trichechus manatus manatus) in Belize and Chetumal Bay,<br />

Mexico. Biological Conservation 95, 67-75.<br />

O'Shea, T.J., Salisbury, C.A.L., 1991. Belize - a last stronghold for manatees in the<br />

Caribbean. Oryx 25, 156-164.<br />

Packard, J. M., Summers R. C., Barnes, L. B. (1985) Variation in visibility bias during<br />

aerial surveys. Journal of Wildlife Management 49: 347-351.<br />

Quinn, G.P., Keough, M.J., 2002. Experimental Design and Data Analysis for Biologists.<br />

Cambridge University Press, Cambridge.<br />

Quintana-Rizzo, E., Reynolds, J.E., 2008. Regional management plan for the West Indian<br />

Manatee (Trichechus manatus), In Caribbean Environment Programme. United Nations<br />

Environment Program, Gosier, Guadeloupe, France.<br />

Reynolds, J.E. III, Wilcox, J. R. 1994. Observations of Florida manatees (Trichechus<br />

manatus latirostris) around selected power plants in winter. Marine Mammal Science<br />

10:163-<strong>17</strong>7.<br />

Scheiner, S.M., Gurevitch, J. eds., 2001. Design and Analysis of Ecological Experiments,<br />

2nd ed. Oxford University Press, Oxford.<br />

Self-Sullivan, C., 2008. Conservation of Antillean Manatees in the Drowned Cays Area of<br />

Belize, In Fisheries and Wildlife. p. 143. Texas A&M.<br />

Self-Sullivan, C. & Mignucci-Giannoni, A. 2008. Trichechus manatus ssp. manatus. In:<br />

IUCN 2011. IUCN Red List of Threatened Species. Version 2011.1.<br />

. Downloaded on 09 November 2011.<br />

Skalski, J.R., Robson, D.S., Simmons, M.A., 1983. Comparative census procedures using<br />

single mark-recapture methods. Ecology 64, 752-760.<br />

Taylor, B.L., Gerrodette, T., 1993. The uses of statistical power in conservation biology:<br />

the vaquita and northern spotted owl. Conservation Biology 7, 489-500.<br />

24


524
<br />

525
<br />

526
<br />

527
<br />

528
<br />

529
<br />

530
<br />

531
<br />

532
<br />

533
<br />

534
<br />

535
<br />

536
<br />

537
<br />

538
<br />

539
<br />

Taylor, B.L., Martinez, M., Gerrodette, T., Barlow, J., Hrovat, Y.N., 2007. Lessons from<br />

monitoring trends in abundance of marine mammals. Marine Mammal Science 23, 157-<br />

<strong>17</strong>5.<br />

Thompson, W.L. ed., 2004. Sampling rare or elusive species. Island Press, Washington.<br />

USFWS, 2001. Florida Manatee Recovery Plan (Trichechus manatus latirostris), Thrid<br />

Revision, ed. U.S.F.W. Service, p. 138 pp. U.S. Fish and Wildlife Service.<br />

Williams, R., Thomas, L., 2009. Cost-effective abundance estimation of rare animals:<br />

Testing performance of small-boat surveys for killer whales in British Columbia. Biological<br />

Conservation 142, 1542-1547.<br />

Wright, I.E., Reynolds J.E. III, Ackerman B. B., Ward, L.I., Weigle, B. L., Szelistowski,<br />

W. A. 2002. Trends in manatee (Trichechus manatus latirostris) counts and habitat use in<br />

Tampa Bay, 1987-1994: implications for conservation. Marine Mammal Science 18:259-<br />

274.<br />

Yoccoz, N.G., Nichols, J.D., Boulinier, T., 2001. Monitoring of biological diversity in<br />

space and time. Trends in Ecology & Evolution 16, 446-453.<br />


<br />

25


Table 1. Results from the GLM relating number of manatees to variables that may influence detection probability (n = 113).<br />

Dependent variable is number of manatees. Independent variables are number of volunteers, total number of observers,<br />

precipitation, cloud cover, sea state, swell height and behavior.<br />

Variables in the Model<br />

df Significance<br />

Wald Chisquare<br />

n = 131<br />

Precipitation (subcategories: dry/light rain)* 0.426 1 NS<br />

Cloud cover (subcategories: clear/scattered/partly cloudy/overcast)* 0.267 3 NS<br />

<br />

<br />

<br />

<br />

<br />

Dependent<br />

variable<br />

Number of manatees<br />

<br />

<br />

Independent variables<br />

Environment<br />

variables<br />

<br />

<br />

<br />

Sea state (subcategories: Beaufort scale <strong>–</strong> 0/1/2/3)* 0.<strong>17</strong>4 3 NS<br />

Glare (subcategores: yes/no)* 0.002 1 NS<br />

Swell Height 0.001 1 NS<br />

Secchi Distance 0.481 1 NS<br />

Sighting Time 0.604 1 NS<br />

Manatee Behavior*<br />

Behavior (subcategories: disturbed, feed, mill, rest, social, travel, other) 2.916 7 NS<br />

*No significant subcategories<br />

<br />

<br />

26


Variables in the Model<br />

df Significance<br />

Wald Chisquare<br />

n = 597<br />

Dependent variable<br />

Number of manatees<br />

Independent variables<br />

Number of volunteers 7.343 1 0.007<br />

Number of volunteers * Year<br />

2001 1.620 7 NS<br />

2002 0.013 1 NS<br />

2003 1.094 1 NS<br />

2004 0.616 1 NS<br />

2005 0.243 1 NS<br />

2006 0.338 1 NS<br />

2007 * * *<br />

Table 2. Generalized linear model (GLM) with a negative binomial distribution and a log-link function to determine<br />

whether the number of manatees sighted was influenced by number of volunteers and year<br />

<br />

<br />

<br />

<br />

27


28


Table 3. Simulation results: Percentage of declines detected (in bold italics) for six different sampling regimens and three different<br />

levels of decline, assuming a perfectly consistent relationship between population index and actual abundance (1000 runs for each<br />

simulation). Scan duration indicates whether the negative binomial distribution parameters were generated from 20 or 30-minute<br />

scans. Number of points indicates the number of points used to create the array of manatee sightings for the simulation models<br />

Parentheses indicates which points were used to generate the negative binomial distribution parameters. > 30% means points with<br />

greater than 30% sighting probability. Number of two-week survey periods indicates the number of survey periods used in the<br />

Sampling Regime<br />

Scan Interval 20-minute scan 30-minute scan<br />

6 8 6 6 8 8<br />

21<br />

(>30 %)<br />

28<br />

(>30 %)<br />

28<br />

(>30 %)<br />

21 points<br />

(all points)<br />

28 points<br />

(>30% )<br />

28 points<br />

(all points)<br />

Num. of two-week survey<br />

periods<br />

Num. of points<br />

(Points used to generate<br />

negative binomial<br />

distribution parameters.)<br />

Results<br />

10% 11 14 9 12 16 13<br />

25% 51 65 33 57 71 60<br />

50% 100 100 99 100 100 100<br />

simulation. We ran our simulation for three potential declines in the number of manatees after 15 years: 10%, 25% and 50%.<br />

These declines are equivalent to a 0.7%, 1.7% and 3.6% decrease in the number of manatees sighted per scan per year.<br />

<br />

<br />

<br />

29


Figure 1. Map of the Drowned Cayes area and scan points<br />

30


a.<br />

b.<br />

c.<br />

Number of sightings<br />

Number of sightings<br />

Number of sightings<br />

100<br />

80<br />

60<br />

40<br />

20<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0<br />

1 2 5 8 9 14 <strong>17</strong> 20 23 26 29<br />


Figure 3. Histogram of number of manatees counted per 20-minute point scan (mean =<br />

0.54, SD = 0.891, n = 960) with negative binomial distribution (solid line).<br />

32


2001 2002 2003 2004 2005 2006 2007<br />

Figure 4. No significant trend in mean number of manatees sighted per 20-minute scan<br />

was detected from 2001-2007 in the Drowned Cayes area of Belize (Likelihood Ratio Chi-<br />

Square = 1.56, df = 1, B = -0.003, p = 0.212 (NS), n = 960). Error bars: +/- 1 SE. Bottom<br />

axis is number of the survey period. Number of manatees per 20-minute scan was the<br />

dependent variable. Independent variable was the time over which the forty-two survey<br />

periods were conducted.<br />

33


a.<br />

b.<br />

c.<br />

Power<br />

Power<br />

Power<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

20 min 6 teams 28 pts<br />

30 min 6 teams 21 pts<br />

30 min 6 teams 28 pts<br />

20 min 8 teams 28 pts<br />

30 min 8 teams 28 pts<br />

<br />

30 min 6 teams 21 pts<br />

2 4 6 8 10 12 14 15<br />

2 4 6 8 10 12 14 15<br />

2 4 6 8 10 12 14 15<br />

Number of years<br />

34<br />

> 30 %<br />

sighting<br />

probability


Figure 5. Power to detect 10% (a), 25% (b), and 50% (c) declines in the number of<br />

manatees over 15 years.<br />

35

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

Saved successfully!

Ooh no, something went wrong!