Download 2014, vol 24 - Department of Biological Sciences

Document related concepts

Bifrenaria wikipedia , lookup

Megafauna wikipedia , lookup

Allometry wikipedia , lookup

River ecosystem wikipedia , lookup

Habitat conservation wikipedia , lookup

Perovskia atriplicifolia wikipedia , lookup

Occupancy–abundance relationship wikipedia , lookup

Tropical Andes wikipedia , lookup

Operation Wallacea wikipedia , lookup

Tropical rainforest wikipedia , lookup

Reconciliation ecology wikipedia , lookup

Theoretical ecology wikipedia , lookup

Latitudinal gradients in species diversity wikipedia , lookup

Habitat wikipedia , lookup

Biological Dynamics of Forest Fragments Project wikipedia , lookup

Transcript
PARTICIPANTS OF THE DARTMOUTH BIOLOGY FSP 2014
FACULTY
RYAN G. CALSBEEK MATTHEW P. AYRES CELIA Y. CHEN
LAB COORDINATOR
CRAIG D. LAYNE
GRADUATE ASSISTANTS
JESSICA V. TROUT-HANEY VIVEK V. VENKATARAMAN
UNDERGRADUATES
ALLEGRA M. CONDIOTTE
MAYA H. DEGROOTE
ANN M. FAGAN
NATHANAEL J. FRIDAY
EMILY R. GOODWIN
LARS-OLAF HOEGER
ALEXA E. KRUPP
AARON J. MONDSHINE
REBECCA C. NOVELLO
CLAIRE B. PENDERGRAST
ELLEN R. PLANE
AMELIA L. RITGER
ZACHARY T. WOOD
1
Dartmouth Studies in Tropical Ecology, Vol. 24 (2014)
Dartmouth College runs an annual 9-10 week ecological field research program in Costa Rica and the Caribbean.
Manuscripts from the research projects in this program have been published in the annual volume “Dartmouth Studies
in Tropical Ecology” since 1989. Copies are held in the Dartmouth library and in Costa Rica at the San Jose office of
the Organization for Tropical Studies (OTS/OET), at the OTS field stations at Palo Verde, Las Cruces and La Selva, at
the Cuerici Biological Station, at the Sirena Station of the Corcovado National Park, and at the Monteverde Biological
Station. On Little Cayman Island, there are copies at the Little Cayman Research Center.
Dartmouth faculty from the Department of Biological Sciences, along with two Ph.D. students from Dartmouth’s
Ecology and Evolutionary Biology graduate program, advise ca. 15 advanced undergraduate students on this program.
The order of authorship on papers is usually alphabetical or haphazard, because all authors contribute equally on
projects. For each paper there is a faculty editor (indicated after the author listing), who takes responsibility for
defining the revisions, and decides on the acceptability of manuscripts for publication. Graduate student Teaching
Assistants are also heavily involved as mentors at every stage, from project design to final manuscript.
We thank the Costa Rican Ministry of the Environment and Energy (MINAE) for permission to conduct research in
Costa Rica’s extraordinary national parks. The Organization for Tropical Studies (OTS/OET) has provided essential
support for our program for over 30 years, taking care of most of our logistical needs in Costa Rica, always to high
standards of quality and reliability. We thank OTS staff at the Palo Verde and La Selva Biological Stations, and at the
Wilson Botanical Garden at Las Cruces, for all their services rendered efficiently, politely and in good spirit. Staff at
the Santa Rosa and Corcovado National Parks have also been gracious in accommodating and assisting us. We thank
Carlos Solano at the Cuerici Biological Station for his depth of knowledge and inspiration. We are grateful to the staff
of the Monteverde Biological Station for access to their facilities, and for making us so comfortable when we arrive
late, dirty, hungry and tired from Santa Rosa.
On Little Cayman Island, the Little Cayman Research Center (LCRC), operated by the Central Caribbean Marine
Institute, is our base for the entire coral reef ecology segment of the program. Expert LCRC staff run the lab, provide
accommodations and food, operate research vessels and take care of SCUBA diving logistics and safety. On the
Dartmouth campus, the Off Campus Programs Office, under the Associate Dean of International and Interdisciplinary
Studies, deals with administration and emergency services and provides an essential lifeline to remote locations in rare
times of need.
We are grateful for the generous financial support of the Biology Foreign Studies Program from Dorothy Hobbs
Kroenlein.
If you have questions about this volume or the program, contact the Biological Sciences Department at Dartmouth
College, Hanover New Hampshire, USA. Currently, the Biology Foreign Studies Program Director is Mstthew Ayres
at [email protected] and the administrative assistant is Sherry L. Finnemore,
[email protected].
Matt Ayres
Hanover NH, USA
21 November 2014
ii
Schedule for Costa Rica
Date
6-Jan
7-Jan
8-Jan
9-Jan
10-Jan
11-Jan
12-Jan
13-Jan
14-Jan
15-Jan
16-Jan
17-Jan
18-Jan
19-Jan
20-Jan
21-Jan
22-Jan
23-Jan
24-Jan
25-Jan
26-Jan
27-Jan
28-Jan
29-Jan
30-Jan
31-Jan
1-Feb
2-Feb
3-Feb
4-Feb
5-Feb
6-Feb
7-Feb
8-Feb
9-Feb
10-Feb
11-Feb
12-Feb
13-Feb
14-Feb
15-Feb
16-Feb
17-Feb
18-Feb
19-Feb
20-Feb
Day Location
Morning
Mon To San Jose
Travel
In San Jose
OTS, InBIO
To Palo Verde
Travel
At PV
Orientation
At PV
FIP-1 (ant-acacia)
At PV
FIP-2
Sun At PV
FIP-2
At PV
SIP-1 plan/proposals
At PV
SIP-1
At PV
SIP-1
At PV
River trip
To Santa Rosa
Travel/walk
At SR
Lec: Mngrv(NL)
Sun To Monteverde Travel
At MV
Orientation
At MV
SIP-2 pilot /props
At MV
SIP-2
At MV
SIP-2
At MV
Writing
At MV
All final mss due
Sun To Cuerici
Travel
At Cuerici
Trip to Paramo
At Cuerici
SIP-3 planning / proposals
At Cuerici
SIP-3
At Cuerici
SIP-3
At Cuerici
Analysis, writing
To La Palma
Exploration at Cuerici
Sun To Sirena
Hike to Sirena
At Sirena
Orientation
At Sirena
SIP-4
At Sirena
SIP-4
To Las Cruces
Hike out of Sirena
At Las Cruces
Orientation. Plant lab.
At Las Cruces
Writing. Botany.
Sun At Las Cruces
Writing. Botany.
At Las Cruces
Writing. Botany.
To La Selva
Travel
At La Selva
Orientation
At La Selva
SIP-5
At La Selva
SIP-5
At La Selva
SIP-5
Sun At La Selva
SIP-5
At La Selva
Analysis. Writing.
At La Selva
Final revisions to all CR mss.
To San Jose
Exploration
To Little Cayman until 10 March
Afternoon
Evening
Travel
Arrive in evening
free: shopping etc.
Group dinner in SJ
Orientation
Lec: Intro CR ecol (RC,Zach)
Research Questions.
Lec: Avian Ecol (JTH, Becca)
Writing clinic, Stat lab (VV)
Data analysis/synthesis
ArthLab (JTH). Peer-review clinic
FIP-1 symposium. Writing.
Lec: DivCoex (RC,Nate ); Peer Review
Writing. FIP-1 ms due
Lec: Primatesl (VV, Lexa). Data /Writ.
FIP-2 seminars. Writing
SIP-1; Peer Review
Writing. FP-2 ms due.
SIP-1/analysis. Revisions.
SIP-1/anal. Revs.
SIP-1 symposium. Writing. Peer Review Writing: SIP-1 ms due
Orientation. Lec:Turtles(RC); VertLab (VV).Field: Sea turtle nesting
Exploration
Field: Sea turtle nesting
Orientation
Writing (revisions)
SIP-2 planning
Lec: Behav (RC,Adam)
SIP-2
Lec: His/Orig (RC,Allegra)
SIP-2
Analysis.Writing (rev.). Peer Review
Analysis/synthesis
SIP-2 symposium
Writing SIP-2 ms due. Bat Jngl
Writing
Exploration
Free
Travel/Orientation
Lec: Coevol 1 (MA, Emily)
Orientation
Writing lab 1 (MA)
SIP-3 pilot
SIP-3 final proposals
SIP-3
Lec: Coevol 2 (HtH, Annie)
SIP-3
Analysis, writing
SIP-3 symposium
Writing SIP-3 ms due
Travel to La Palma
Sirena preparation
Hike to Sirena
Natural history reports from hike
SIP-4 plan
Lec: Social insects (VV, Claire)
SIP-4
Lec: Plant-Herb (MA, Aaron)
SIP-4. Adios Hannah.
Analysis
Travel to Las Cruces
Analysis. Writing
Analysis. Writing
SIP-4 symposium.
Writing. Botany.
Writing lab 2 (MA)
Writing. Botany.
Writing
Writing. Botany practicum.
Discussion: Why Science?
Travel
Analysis. Writing
SIP-5 plan/pilot
Lec: Aquatic Ecology (JTH, Maya)
SIP-5
Writing. Night walk.
Agroecology field trip
Lec: Ecosystems (MA, Ellen)
SIP-5
Paper (Lars). Analysis. Writing. Walk.
Analysis. Writing
Paper (Amelia). Analysis. Writing. Walk.
Analysis. Writing
SIP-5 symposium.
Travel to SJ
Lec: Cons. Biology (MA)
Travel to SJ
Free
FIP = Fa cul ty i ni ti a ted project; SIP = Student i ni ti a ted project
iii
LITTLE CAYMAN 2014 SCHEDULE
Date
20Feb
21Feb
Arrive from
CR
Orientation
Morning
Afternoon
Evening
Main orientation and safety
information – Rob, Celia,
Lowell, Greg
Get BCD and regulators
assigned from Lowell
Discussion: program to
date and expectations for
LC segment.
Lecture: Coral biology
lecture & Reef
morphology (CC)
SCUBA –shore dive
at Cumber’s Cave
(check dive and
snorkel)
Critique: Coral biology
(Lars-Olaf Hoeger)
Critique: Coral Disease
(Zach Wood)
Discussion: Natural
History, Queen
Conch
22Feb
General
natural
history
Lecture: Algae and Sea &
lab (CC)
Preston Bay - Snorkel to
see invertebrates and
collect algae followed
Queen Conch Project
Critique: Seagrass
beds (Emily
Goodwin)
Discussion: Natural
History
23Feb
Queen
Conch
project
Project 1
starts
Queen Conch Project
Queen Conch Data
collection
Work on Queen Conch data
analysis and write up
9am Queen Conch
Project Presentation
Lecture: Invertebrates
(CC)
Critique: Invertebrates
(Aaron Mondshine)
Project 1 exploration
Critique: Sponges
(Rebecca Novello)
Project 1 idea discussion
25Feb
SCUBA
Project 1 brainstorm
26Feb
Project 1 Proposal
Presentation
Lecture: Reef Nekton
(CC)
Critique: Fish ecology
(Ellen Plane)
Lecture: Fish behavior
(VVV)
Project 1
Project 1
Zooplankton lab
(JTH)
Finalize project 1 idea,
design, and group members
Lecture: Zooplankton
(JTH) Critique:
Zooplankton (Allegra
Condiotte)
Critique: Herbivory
(Amelia Ritger 5PM)
24Feb
27Feb
Project 1
iv
Critique: Fish Behavior
(Nathanael Friday)
Critique: Fish Biology
(Annie Fagan)
28Feb
Date
1-Mar
Dia libre
(OFF)
2-Mar
3-Mar
Project 1
SCUBA
2:30 pm Skype with
John Bothwell,
Senior Research
Officer, Cayman
Environment Centre
Morning
Beach Cleanup
11AM
Afternoon
Relax/Cookout
**Project 1
PRESENTATIONS**
4 PM Project 1 – 1st
Draft due
Lecture: Lionfish
and coral nursery,
Katie Lohr
Lecture: : Coral reefs and
climate change (CC)
Project 1 – data analysis,
write methods, stats help
RR: Karaoke for those
interested
Evening
Lecture Conservation and
Management (CC)
Critique: Food webs
(Adam Schneider
Critique: Coral Reef
Bleaching (Claire
Pendergast)
Project 2 brainstorming
Project 1 editing and
revision
Finalize project 2 idea,
design, and group members
Project 2 brainstorming
and pilot
Project 2
1pm - Project 2 idea
discussion
4-Mar
8 AM Proposal 2
Presentation
Project 2
4 PM Project 2
proposal DUE
Critique: Fish and Climate
Change (Maya DeGroote)
SCUBA Night Dive
5-Mar
Project 2
Project 2
Lecture: Mangroves (CC)
Discussion: Graduate
School/Career
6-Mar
Project 2
Project 2
7-Mar
SCUBA
Project 2
**Project 2 1st draft
DUE @ 12 pm**
Critique: Mangroves (Lexi
Krupp)
Project 2 – writing and
analysis
**Project 2 Presentation**
Project 2 refine drafts
Project 2
starts
8-Mar
Mardi Gras parade
Project 2 refine drafts
9-Mar
Breakfast Brunch
Clean up science
equipment/field sites
Project 2 final draft due
Revisions & copy
editing of all projects
Individual meetings
10Mar
Depart LC
8:20 or 10:00
am
v
RR: Karaoke for those
interested
Project 2 refine drafts
Cookout and Bonfire
(pending wind)
Individual meetings
Clean up & pack
PAPERS FOR STUDENT PRESENTATIONS: COSTA RICA
Site
Lecture
Student
PV
Intro CR
Ecol
Zach
PV
Primate
Ecol
Lexi
PV
Avian Ecol
Becca
PV
Behav
Adam
PV
MV
Diversity &
Coexistenc
e
History &
Origins
Nate
Allegra
Cuer
Coevol 1
Emily
Cuer
Coevol 2
Annie
Corc
Social
Insects
Claire
Corc
PlantHerbiv
Aaron
LaSelva
Aquatic
Ecol
Maya
LaSelva
Ecosystems
Ellen
LaSelva
Wildcard
Lars
LaSelva
Cons.
Biology
Amelia
Student Paper
McCain, C. 2009. Vertebrate range sizes indicate that
mountains may be 'higher' in the tropics. Ecology
Letters 12:550-560. (*See also: Janzen, D. H. 1967.
Why mountain passes are higher in the tropics.
American Naturalist 101:230-243.)
Fedigan, L. M., S. D. Carnegie, and K. M. Jack. 2008.
Predictors of reproductive success in female whitefaced capuchins (Cebus capucinus). Amer. J. Physical
Anthropology 137:82-90.
Botero, C. A., R. J. Rossman, L. M. Caro, L. M.
Stenzler, I. J. Lovette, S. R. de Kort, and S. L.
Vehrencamp. 2009. Syllable type consistency is related
to age, social status and reproductive success in the
tropical mockingbird. Animal Behaviour 77:701-706.
Irwin D.E. et al. 2001. Speciation in a ring. Nature 409,
333-337.
Molino, J.-F. and D. Sabatier. 2001. Tree diversity in
tropical rain forests: a validation of the intermediate
disturbance hypothesis. Science 294:1702-1704.
Janzen, D. H. 1981. Neotropical anachronisms: the
fruits the gomphotheres ate. Science 215:19-27.
Ramirez, S. R., T. Eltz, M. K. Fujiwara, G. Gerlach, B.
Goldman- Huertas, N. D. Tsutsui, and N. E. Pierce.
2011. Asynchronous diversification in a specialized
plant-pollinator mutualism. Science 333:1742-1746.
Herre, E. 1993. Population-structure and the evolution
of virulence in nematode parasites of fig wasps. Science
259:1442-1445.
Wray, M. K., H. R. Mattila, and T. D. Seeley. 2011.
Collective personalities in honeybee colonies are linked
to colony fitness. Animal Behaviour 81:559-568.
Kursar T. A., et al. 2009. The evolution of antiherbivore
defenses and their contribution to species coexistence in
the tropical tree genus Inga. PNAS 106:18073-18078.
Pringle, C., T. Hamazaki. 1998. The role of omnivory
in a neotropical stream: Separating diurnal and
nocturnal effects. Ecology 79:269-280.
Higgins S. I., S. Scheiter. 2012. Atmospheric CO2
forces abrupt vegetation shifts locally, but not globally.
Nature 488:209-212.
Caldera, E. J., C. R. Currie. 2012. The population
structure of antibiotic-producing bacterial symbionts of
Apterostigma dentigerum ants: Impacts of coevolution
and multipartite symbiosis. American Naturalist
180:604-617.
Becker, C. G., K. R. Zamudio. 2011. Tropical
amphibian populations experience higher disease risk in
natural habitats. PNAS 108:9893- 9898.
vi
Staff
Ryan
Vivek
Jess
Ryan
Ryan
Ryan
Matt
Hannah
Vivek
Matt
Jess
Matt
Matt
Matt
PAPERS FOR STUDENT PRESENTATIONS: LITTLE CAYMAN
Student
Allegra Condiotte
Lecture
Zooplankton
Maya DeGroote
Fish biology
Annie Fagan
Fish biology
Nathanael Friday
Fish behavior
Emily Goodwin
Seagrass beds
Lars-Olaf Hoeger
Coral biology
Alexa Krupp
Aaron Mondshine
Becca Novello
Claire Pendergast
Ellen Plane
Mangroves
Invertebrates
Sponges
Coral bleaching
Fish ecology
Amelia Ritger
Herbivory
Adam Schneider
Food webs
Zach Wood
Coral disease
Paper
Heidelberg, K. B., K. P. Sebens, and J. E. Purcell. 2004. Composition and sources of near ree
along with implications for coral feeding. Coral Reefs 23:263-276.
Miller, G.M., S. Watson, S., J.M. Donelson, M.I. McCormick, P.L.
Munday. Parental environment mediates impacts of increased carbon
dioxide on a coral reef fish. Nature Climate Change 2: 858-861.
Gerlach, G., J. Atema, M. J. Kingsford, K. P. Black, and V. MillerSims. 2007. Smelling home can prevent dispersal of reef fish larvae.
Proceedings of the National Academy of Sciences of the United
States of America 104:858-863.
Grutter, A. S., J. M. Murphy, and J. H. Choat. 2003. Cleaner fish
drives local fish diversity on coral reefs. Current Biology 13:64-67.
Tewfik, A., J. Rasmussen, and K. S. McCann. 2005. Anthropogenic
enrichment alters a marine benthic food web. Ecology 86:2726-2736.
Wild, C., M. Huettel, A. Klueter, S. G. Kremb, M. Y. M. Rasheed,
and B. B. Jorgensen. 2004. Coral mucus functions as an energy
carrier and particle trap in the reef ecosystem. Nature 428:66-70.
Mumby, P. J., A. J. Edwards, J. E. Arias-Gonzalez, K. C. Lindeman,
P. G. Blackwell, A. Gall, M. I. Gorczynska, A. R. Harborne, C. L.
Pescod, H. Renken, C. C. C. Wabnitz, and G. Llewellyn. 2004.
Mangroves enhance the biomass of coral reef fish communities in the
Caribbean. Nature 427:533-536.
Carpenter, R. C. and P. J. Edmunds. 2006. Local and regional scale
recovery of Diadema promotes recruitment of scleractinian corals.
Ecology Letters 9:268-277.
De Goeij, J.M., D. vab Oevelen, M.J.A. Verimeij, R. Osinga, J.J.
Middelburg, A.F.P.M. de Goeij, W. Admiraal. 2013. Surviving in a
marine desert: the sponge loop retains resources within coral reefs.
Science 342: 108-110.
Grottoli, A.G., L.J. Rogriguez and J.E. Palardy. 2006. Heterotrophic
plasticity and resilience in bleached corals. Nature 440: 1186-1189.
McMahon, K.W., M.L. Berumen, and S.R. Thorrold. Linking habitat
mosaics and connectivity in a coral reef seascape. Proceedings of the
National Academy of Sciences 109: 15372-15376.
Dixson, D. L. and M. E. Hay. 2012. Corals Chemically Cue
Mutualistic Fishes to Remove Competing Seaweeds. Science
338:804-807.
Mumby, P. J., C. P. Dahlgren, A. R. Harborne, C. V. Kappel, F.
Micheli, D. R. Brumbaugh, K. E. Holmes, J. M. Mendes, K. Broad, J.
N. Sanchirico, K. Buch, S. Box, R. W. Stoffle, and A. B. Gill. 2006.
Fishing, trophic cascades, and the process of grazing on coral reefs.
Science 311:98-101.
Patterson, K. L., J. W. Porter, K. E. Ritchie, S. W. Polson, E.
Mueller, E. C. Peters, D. L. Santavy, and G. W. Smiths. 2002. The
etiology of white pox, a lethal disease of the Caribbean elkhorn coral,
Acropora palmata. Proceedings of the National Academy of Sciences
of the United States of America 99:8725-8730.
vii
MAPS: COSTA RICA AND LITTLE CAYMAN ISLAND
http://www.nationsonline.org/oneworld/map/costa-rica-map.htm
http://www.southerncrossc
lub.com/cayman-islands/
viii
Table of Contents
Participants of the 2014 FSP
i
Note from Professor Ayres
ii
Costa Rica 2014 Schedule
iii
Little Cayman 2014 Schedule
iv-v
Papers for Student Presentations in Costa Rica
vi
Papers for Student Presentations in Little Cayman
vii
Maps
viii
Palo Verde
The ant, Pseudomyrmex spinicola, responds to variation in reward
level by the bullhorn acacia, Acacia collinsii. Maya H. DeGroote,
Aaron J. Mondshine, Rebecca C. Novello, and Amelia L. Ritger
1
Forgetful Flavicornis: acacia ants do not react more vigorously to
repeated attacks. Allegra M. Condiotte, Lars-Olaf Höger, Alexa E.
Krupp, Ellen R. Plane, and Adam N. Schneider
4
Explaining the distribution of tropical Myrmeleontidae antlions:
contributions of habitat, prey availability, and feeding dynamics.
Ann M. Fagan, Nathanael J. Friday, Emily R. Goodwin, Claire B.
Pendergrast, and Zachary T. Wood
6
Explaining the distribution of wetland macroinvertebrates:
contributions of habitat, prey availability and feeding dynamics.
Allegra M. Condiotte, Nathanael J. Friday, Aaron J. Mondshine, Adam
N. Schneider, and Zachary T. Wood
9
Behavioral thermoregulation based on coloration in tropical
butterflies. Maya H. DeGroote, Alexa E. Krupp, Rebecca C. Novello,
Claire B. Pendergrast, and Amelia L. Ritger
12
Factors affecting vigilance behavior in the primates, Alouatta palliata
and Cebus capucinus. Ann M. Fagan, Emily R. Goodwin, Lars-Olaf
16
ix
Höger, and Ellen R. Plane
Ontogeny and predatory efficiency in Centruroides margaritatus. Ann
M. Fagan, Emily R. Goodwin, Aaron J. Mondshine, Ellen R. Plane, and
Amelia L. Ritger
Wetland macrophyte diversity affects fish abundance, morphology,
and species richness. Allegra M. Condiotte, Maya H. DeGroote, Alexa
E. Krupp, Adam N. Schneider, and Zachary T. Wood
Patterns of nutrient distribution between soil and aboveground plant
biomass in tropical forests do not hold at the understory level.
Nathanael J. Friday, Lars-Olaf Höger, Rebecca C. Novello, and Claire B.
Pendergrast
19
22
25
Monteverde
Orchidaceae diversity in tropical primary and secondary forest. Ann
M. Fagan, Alexa E. Krupp, Aaron J. Mondshine and Amelia L. Ritger
28
Thermoregulatory cost of dominance behaviors in hummingbirds.
Allegra M. Condiotte, Emily R. Goodwin, Rebecca C. Novello, and
Adam N. Schneider
33
Optimizing leaf inclination angles: water shedding and
photosynthesis in cloud forest understory plants. Maya H. DeGroote,
Claire B. Pendergrast, and Ellen R. Plane
36
Forest structure and epiphyte richness. Nathanael J. Friday, Lars-Olaf
Höger, and Zachary T. Wood
40
Cuerici
Differential mobbing in tropical highland birds. Ann M. Fagan, Emily
R. Goodwin, and Aaron J. Mondshine
43
Diversity or dominance: synchronous masting as a reproductive
strategy in Quercus costaricensis montane forests. Maya H. DeGroote
and Claire B. Pendergrast
47
Insectivorous bats alter foraging decisions based on risk of visual
predation. Alexa E. Krupp, Rebecca C. Novello, and Ellen R. Plane
52
x
Spatial patterning in the abundance of floral mites phoretic on
hummingbirds. Lars-Olaf Höger and Zachary T. Wood
56
Female mate choice and male lekking behavior in Drosophila. Amelia
L. Ritger
61
Size distribution and aggression in rainbow trout, Oncorhynchus
mykiss. Allegra M. Condiotte, Nathanael J. Friday, and Adam N.
Schneider
66
Corcovado
Aquatic community composition of the Río Claro estuary. Allegra M.
Condiotte, Nathanael J. Friday, Lars-Olaf Höger, Aaron J. Mondshine,
Ellen R. Plane, and Amelia L. Ritger
71
Constraints on resource transport in Atta colombica. Maya H.
DeGroote, Ann M. Fagan, Emily R. Goodwin, Adam N. Schneider, and
Zachary T. Wood
77
The spatial dispersion of immature and adult tropical cicadas. Alexa
E. Krupp, Rebecca C. Novello, and Claire B. Pendergrast
82
La Selva
Spatial distribution of Lycosidae spiders within and across habitats.
Alexa E. Krupp, Aaron J. Mondshine, Claire B. Pendergrast, Amelia L.
Ritger, and Adam N. Schneider
87
The rich get redder: delayed greening in Welfia regia palms. Allegra
M. Condiotte, Ann M. Fagan, Nathanael J. Friday, Ellen R. Plane, and
Zachary T. Wood
91
Log preference in Lepidophyma flavimaculatum: shelter, hunting
ground, or both? Maya H. DeGroote, Lars-Olaf Höger, and Rebecca C.
Novello
95
Behavioral thermoregulation in facultatively poikilothermic sloths
(Bradypus Variegatus and Choloepus hoffmanni). Emily R. Goodwin
99
xi
Little Cayman
Patterns in coral assemblages of a patch reef system. Maya H.
DeGroote, Emily R. Goodwin, Aaron J. Mondshine, and Ellen R. Plane
105
Grazing and epibionts in marine turtle grass beds. Allegra M.
Condiotte, Nathanael J. Friday, Lars-Olaf Höger, and Zachary T. Wood
110
Burrow occupation and movement in Echinometra lucunter. Ann M.
Fagan, Rebecca C. Novello and Claire B. Pendergrast
114
A. palmata colony health and size affect fish abundance, not species
diversity. Alexa E. Krupp, Amelia L. Ritger, and Adam N. Schneider
120
The role of heterotypic schooling in foraging behavior of parrotfish
Emily R. Goodwin and Aaron J. Mondshine
124
Salt pond minnows acclimatize to salinity. Rebecca C. Novello,
Amelia L. Ritger, and Zachary T. Wood
128
Energy allocation varies throughout growth and between sexes of
Pterois volitans. Maya H. DeGroote and Alexa E. Krupp
132
Impact of protected areas on spiny lobster (Panulirus argus)
distribution and population structure. Allegra M. Condiotte,
Nathanael J. Friday, and Claire B. Pendergrast
136
My anemone, my rules: niche partitioning in anemone communities.
Lars-Olaf Höger, Ellen R. Plane, and Adam N. Schneider
143
xii
Dartmouth Studies in Tropical Ecology 2014
THE ANT, PSEUDOMYRMEX SPINICOLA, RESPONDS TO VARIATION IN REWARD LEVEL BY
THE BULLHORN ACACIA, ACACIA COLLINSII
MAYA H. DEGROOTE, AARON J. MONDSHINE, REBECCA C. NOVELLO AND AMELIA L. RITGER
Project Designer: Jessica Trout-Haney
Faculty Editor: Ryan Calsbeek
Abstract: In the coevolution of plants and herbivores, plants must balance energy among growth, reproduction, and
defense. The mutualism between the bullhorn acacia tree, Acacia collinsii, and its protective ant, Pseudomyrmex
spinicola, provides a useful model to examine plant investment in defense systems. Given the acacia's significant
energetic input in producing carbohydrate- and protein-rich ant rewards in the form of extrafloral nectaries and
Beltian Bodies, we hypothesized that ants would be sensitive to variations in reward level. To test our hypothesis,
we measured ant movement and ant abundance after applying treatments of varying artificial reward to three acacia
trees. Ants were not significantly affected by changes in reward level. Although our results were non-significant
they suggest that ants may be sensitive to variation in nutritional rewards by host plants. This would indicate that
acacia reward levels are the product of an energetic tradeoff driven by plant-herbivore interactions.
Key words: mutualism, plant defense theory, Pseudomyrmex spinicola, rewards, Vachellia collinsii
INTRODUCTION
In the evolutionary arms race between plants and
herbivores, plants must carefully balance energetic
inputs among growth, reproduction and defense
(Frederickson et al 2012). Plants can adopt a
variety of defense strategies including chemical
defenses, such as toxins and digestibility reducers,
and mechanical defenses, such as thorns and
trichomes. Additionally, many plants recruit other
organisms for defense against herbivores in the
form of mutualisms (Bronstein 1998; Song and
Feldman 2013).
One such mutualism occurs in the tropical dry
forest of Central America between the bullhorn
acacia tree, Acacia collinsii, and the ant,
Pseudomyrmex spinicola. The tree provides ants
with nutrition in the form of carbohydrate- and
protein-rich extrafloral nectaries (EFNs) and
Beltian Bodies, as well as shelter in the form of
thorns on its stems (Wood 2005). Typically, the
EFNs provide the ants with 3-11% sugar solution
(Heil 2004). In return, ants provide the tree with
protection against herbivorous predators (Janzen
1966; Coley and Barone 1996) by quickly
responding to physical disturbances on the plant
and attacking the source of disturbance (Young et
al 2008). Since evolution favors those plants that
are able to minimize cost and maximize defense,
the acacia tree should expend energy to provide
ants with rewards only if the energy investment
truly increases protection by the ants (Mazancourt
et al 2001). If ants were insensitive to reward
level, we would expect to see the acacia plant
minimizing energy investment into reward.
However, since acacias produce various nutrientand energy-rich rewards, we hypothesized that the
ants would be sensitive to variation in reward,
adjusting their activity based on reward level.
To test this hypothesis, we simulated varying
levels of reward by acacia plants and measured ant
sensitivity to this variation. We predicted that
there would be a positive relationship between
reward level and both ant abundance on a leaf and
ant movement to and from a leaf.
METHODS
We haphazardly selected three acacias of similar
size and ant species in dry shady areas of an acacia
grove in Palo Verde National Park, Costa Rica on
January 10, 2014 at 0830. For each tree, we chose
three leaves of comparable height on the main
stem to receive one of three treatments: (1) high
reward, (2) a control, or (3) no reward. For the
high reward treatment we applied thirty droplets of
a 70% sugar solution on leaflets near the base of
the leaf and tapped the base several times. We
used a 70% solution in order to simulate a higher
potency reward than that naturally produced by
EFNs. For the control treatment, we applied thirty
droplets of tap water and tapped the base of the
leaf in an identical manner. For the no reward
treatment, we applied thirty droplets of tap water
and clipped the EFNs located at the base of the
each selected leaf. Tapping was included in the
1
Palo Verde
control and high reward treatments to standardize
for any potential disturbance response due to
clipping the EFNs in the no reward treatment.
Before we applied the treatments, we recorded
a baseline count of ant movement, which we
defined as the number of ants that passed over the
EFN in a two-minute interval. We disregarded
directionality for all measurements of ant
movement. After treating each leaf, we waited one
hour to eliminate any immediate ant response to
herbivory. After one hour we measured ant
movement and ant abundance on each
experimental leaf. We defined ant abundance as
the number of ants present on a leaf. We used both
change in ant movement and abundance after
treatment as indicators of ant sensitivity to reward.
We performed ANOVAs to test differences in
ant abundance and change in ant movement across
treatments. We also performed a two-tailed
matched-pair t-test to determine whether any
treatments resulted in a significant change in ant
movement. All analyses were performed using
JMP® 10 statistical software.
RESULTS
We found a non-significant difference in ant
abundance across reward treatments (ANOVA,
F2,8 = 1.93, P = 0.23). Ant abundance was greatest
in the high reward treatment and lowest in the no
reward treatment. We found no significant
difference in change in ant movement across
treatments (ANOVA, Fig. 2, F2,8 = 2.14, P = 0.20).
None of the treatments resulted in a significant
change in movement (two-tailed matched-pair t-
2
Figure 2. There was no significant difference in change
in ant movement across treatments (ANOVA, F2,8 =
2.14, P = 0.20). Ant movement was calculated as the
number of ants passing over the extra-floral nectary at
the base of the specified leaf in either direction.
test, high reward: t2 = -1.56, P = 0.26; control: t2 =
1.89, P = 0.20; no reward: t2 = -1.56, P = 0.86).
DISCUSSION
We found that ants did not significantly change
their movement based on host rewards. However,
there was a trend towards decreased ant movement
in response to high reward (Fig. 2). Decreased ant
movement on high reward treatments could be due
to increased sugar concentration compelling ants
to remain on a leaf. Ant abundance might
therefore be a better proxy for ant sensitivity than
ant movement. Furthermore, change in ant
movement in the no reward treatment varied
greatly, suggesting there may be other
consequences to mechanically removing of EFNs
from the stem (e.g., a chemical response to
clipping).
We found a non-significant trend in ant
abundance that indicates ants are potentially
sensitive to reward (Fig. 1). Ant abundance was
greatest on leaves with high reward and lowest on
leaves with no reward. If this trend in ant
abundance is valid and ants are sensitive to
alterations in reward levels, then the bullhorn
acacia tree may be caught in an energetic tradeoff.
This would suggest that acacias are producing an
intermediate level of nutritional reward that
satisfies the ants’ needs and lies within the plant’s
energy budget. However, since our findings were
non-significant, it is also possible that ants are not
actually sensitive to variations in reward, or that
they are only sensitive to repeated applications of
reward for longer periods of time. Long-term
Dartmouth Studies in Tropical Ecology 2014
experimental tests of ant sensitivity remain
important in better understanding how energetic
tradeoffs drive patterns in plant-herbivore
interactions.
ACKNOWLEDGEMENTS
We would like to thank Jess Trout-Haney for her
insight on our project. We would also like to thank
LITERATURE CITED
Coley, P.D. and J.A. Barone. 1996. Herbivory and
plant defenses in tropical forests. Annual
Review of Ecology and Systematics 27: 305335.
Bronstein, J.L. 1998. The contribution of ant-plant
protection studies to our understanding of
mutualism. Biotropica 30: 150-161.
de Mazancourt, C., M. Loreau, and U. Dieckmann.
2001. Can the evolution of plant defense lead
to plant herbivore mutualism. The American
Naturalist 158: 109-123.
Frederickson, M.E., A. Ravenscraft, G.A. Miller,
L.M. Arcila Hernandez, G. Booth, and N.E.
Pierce. 2012. The direct and ecological costs
of an ant-plant symbiosis. The American
Naturalist 179: 768-778.
Heil, M., B. Baumann, R. Krüger, and K.E.
Linsenmair. 2004. Main nutrient compounds
in food bodies of Mexican Acacia-ant plants.
Chemoecology 14: 45-52.
Heil, M., M. Gonzalez-Teuber, L.W. Clement, S.
the 2014 Dartmouth Biology FSP group for peer
reviewing our paper.
AUTHOR CONTRIBUTIONS
All authors contributed equally in the experiment
and writing process.
Kautz, M. Verhaagh, and J.C.S Bueno. 2009.
Divergent investment strategies of acacia
myrmecophytes and the coexistence of
mutualists and exploiters. PNAS 106: 1809118096.
Janzen, D.H. 1966. Coevolution of mutualism
between ants and acacias in Central America.
Evolution 20: 249-275.
Song, Z., and M.W. Feldman. 2013. Plant-animal
mutualism in biological markets: evolutionary
and ecological dynamics driven by nonheritable phenotypic variance. Theoretical
Population Biology 88: 20-30.
Young, H., L. Fedigan, and J. Addicott. 2008.
Look before Leaping: Foraging Selectivity on
Capuchin Monkeys on Acacia Trees in Costa
Rica. Oecologia 155: 85-92.
Wood, W.F. 2005. A comparison of mandibular
gland volatiles from ants of the bull horn
acacia, acacia collinsii. Biochemical
Systematics and Ecology 33: 651-658.
3
Palo Verde
FORGETFUL FLAVICORNIS: ACACIA ANTS DO NOT REACT MORE VIGOROUSLY TO
REPEATED ATTACKS
ALLEGRA M. CONDIOTTE, LARS-OLAF HÖGER, ALEXA E. KRUPP, ELLEN R. PLANE, AND
ADAM N. SCHNEIDER
Project Designer: Vivek Venkataraman
Faculty Editor: Ryan G. Calsbeek
Abstract: Concentrated herbivore damage to one area of a plant generally has more deleterious effects on plant
fitness than a comparable amount of damage dispersed throughout the plant. Using the ant-acacia defense mutualism
as a model, we address the question of whether Pseudomyrex flavicornis will respond more vigorously to a second
attack at the site of recent damage than to an attack at a previously undamaged site. We simulated herbivory on
acacia trees to determine whether there was higher ant recruitment to branches receiving multiple attacks. We found
no significant relationship between repetitive disturbance and the strength of the ant response, suggesting that acacia
ants have not evolved to protect their host plant more vigorously against concentrated herbivory.
Key words: Acacia collinsii, herbivory, mutualism, Pseudomyrmex flavicornis
INTRODUCTION
Plant fitness is often reduced more by
concentrated herbivory to one area than by an
equivalent amount of damage dispersed
throughout the plant (Mauricio et. al 1993). It
should therefore be advantageous for plants to
avoid repeated damage to the same area.
In the ant-acacia mutualism, ants defend the
tree against competition and herbivory and are
rewarded with shelter and nourishment provided
by the tree (Cronin 1996). Mechanical disturbance
to Acacia collinsii induces a chemical response
that recruits Pseudomyrex flavicornis to attack
potential herbivores (Agrawal 1998), thus
discouraging herbivory on A. collinsii.
Plants are capable of modifying their
defensive strategies in response to changing
external stimuli: in plants with ant mutualists,
many different cues, both mechanical and
chemical, result in ant recruitment and defense
(Agrawal and Rutter 1998). In the ant-acacia
relationship, ants could reduce the severity of a
concentrated attack by increasing the number of
ants responding to a recent attack site upon
subsequent disturbances.
The goal of this study was to determine
whether P. flavicornis protect against concentrated
damage to their host plant by responding more
intensely to repeat attacks. We predicted that P.
flavicornis would respond more vigorously to a
second attack at a site of recent damage than to an
attack at a previously undamaged site.
4
METHODS
We measured P. flavicornis activity on A. collinsii
trees in Palo Verde National Park, Costa Rica on
January 10, 2014 from 08:30 to 10:30.
We haphazardly chose five sites consisting of
two adjacent trees. We chose one branch with
similar orientation, height, and sun exposure on
each tree and randomly designated one as control
and one as experimental. We recorded baseline
activity levels at both branches by counting the
number of ants that passed a designated point 5-10
cm from the distal end of the branch in either
direction for one minute.
For experimental branches in the first trial, we
clipped five terminal leaves with scissors and then
tapped the branch with the scissors close to the
end of the branch five times to create further
mechanical disturbance. At the control branches,
we made five snipping noises with scissors next to
the end of the branch to control for sound
disturbance. Ten seconds after each treatment, we
again recorded ant activity. We waited one hour to
allow ant activity to return to pre-disturbance
levels. Upon return, we recorded a second baseline
activity level. For the second trial we carried out
an identical attack on both control and
experimental branches and measured ant activity
as above.
Statistical Analyses and Modeling
We conducted a matched pairs t-test on the first
and second baseline activity levels of experimental
branches. We calculated the effect of prior
Dartmouth Studies in Tropical Ecology 2014
disturbance by comparing the percent change in
ant response (post-attack minus second baseline)
between control and treatment (previously
attacked) trees. We used JMP11 software (SAS
Institute, Cary, NC) to perform a 2-tailed t-test
comparing the difference in means between these
two groups.
RESULTS
Simulated herbivory produced a significant
increase in ant activity (t9 = -3.33, P = 0.01) with a
mean response of 9.3 ± 1.4 (±1SE) ants moving
to and from the site of disturbance. There was no
Figure 1: No significant difference in ant response
was observed between control and experimental
treatments.
significant difference between baseline ant counts
on experimental branches in the first trial
compared to the second (t4 = 0.95, P = 0.40).
The percent change in ant response between
branches that were previously attacked and control
branches was not significant (t9 = 1.48, P = 0.18;
Fig. 1). The mean number of ants responding to
attack during the second trial was 21±7.4 (±1SE)
for previously attacked branches and 26±10.3
LITERATURE CITED
Agrawal, A.A. 1998. Leaf damage and associated
cues induce aggressive ant recruitment in a
neotropical ant-plant. Ecology 79: 2100-2112.
Agrawal, A.A. and Rutter, M.T. 1998. Dynamic
anti-herbivore defense in ant-plants: the role
of induced responses. Oikos 83: 227-236.
Conin, G. 1996. Between-species and temporal
variation in Acacia-ant-herbivore interactions.
(±1SE) for control branches.
DISCUSSION
The average number of ants responding to attacks
administered during the second trial did not differ
between first attack and repeat attack. Ants did not
have an elevated response in previously attacked
areas, suggesting that the A. collinsii and P.
flavicornis association does not have a “memory”
for concentrated herbivory.
Although the attacks clearly stimulated an
increase in ant activity, after the hour wait
between trials, activity had reverted to pre-attack
levels. Therefore the second attack was not
affected by lingering activity induced by the first
attack. Since we observed increased activity
immediately after an attack, future studies may
benefit from a reduced wait time between the
simulated attack and the monitoring of defensive
response.
Our results suggest that in the ant-acacia
mutualism, the benefits of the ants mitigating a
concentrated attack do not outweigh the cost of
increased defense. It is, however, unclear for
which partner the energetic costs outweigh the
benefits. It is also possible that the ants interpret a
branch with repeated damage as a compromised
resource, and that a low-resource branch is no
longer worth the extra energy expense of increased
defense. These findings emphasize the need for
further study on the distribution of energy costs
between mutualistic partners.
ACKNOWLEDGEMENTS
We would like to thank Vivek Venkataraman for
his guidance in planning and executing this study.
AUTHOR CONTRIBUTIONS
All authors contributed equally in the experiment
and writing process.
Biotropica 30: 135-139.
Janzen, D.H. 1966. Coevolution of mutualism
between ants and acacias in Central America.
Evolution 20: 249-275.
Mauricio, R., Bowers, M.D., and Bazzaz, F.A.
1993. Pattern of leaf damage affects fitness of
the annual plant Raphanus sativus
(Brassicaceae). Ecology 74: 2066-2071.
5
Palo Verde
EXPLAINING THE DISTRIBUTION OF TROPICAL MYRMELEONTIDAE ANTLIONS:
CONTRIBUTIONS OF HABITAT, PREY AVAILABILITY, AND FEEDING DYNAMICS
ANN M. FAGAN, NATHANAEL J. FRIDAY, EMILY R. GOODWIN,
CLAIRE B. PENDERGRAST, AND ZACHARY T. WOOD
Project Designer: Ryan Calsbeek
Faculty Editor: Ryan G. Calsbeek
Abstract: Predator distributions are generally limited by prey availability and habitat suitability. We compared the
importance of prey preference and substrate type in determining the distribution of the Myrmeleontidae antlion in
Palo Verde National Park. We hypothesized that substrate type was the primary driver of trap site selection, trap
size, and predation success for larval antlions. To test this hypothesis, we censused antlion traps across different
substrate types and experimentally assessed antlion substrate preference. We tested antlion capture success across a
range of trap sizes in sandy substrate to determine trap size’s impact on prey capture success. Antlions significantly
preferred sandy substrate over hard and loose soil. Antlion capture success increased with trap size, while prey
species had no effect. This indicates that antlion predation is more strongly limited by suitable substrate for efficient
trap construction than by prey type or abundance. Antlions and other predators who modify their environment to
catch their prey are strongly limited by habitat conditions.
Key words: Myrmeleontidae, species distributions, predator-prey dynamics, burrower-soil interactions
INTRODUCTION
Species distributions are linked to resource
availability and community interactions, such as
predation, prey availability, and habitat type (Gatti
and Farji-Brener 2002). Predatory organisms aim
to optimize food intake through a variety of
predation strategies including ambush and trapbuilding. Organisms whose predation strategies
necessitate a specific habitat type may be more
limited by habitat availability than prey
availability.
Bull Horn Acacia (Vachellia collinsii; Seigler
and Ebinger 2005) are hosts in a mutualism
(Janzen 1966) with red acacia ants
(Pseudomyrmex spinicola). One of the facets of
the ant-acacia mutualism is the eradication of
competing vegetation under the acacia trees by the
ants, leaving the soil dry and hard. A common
predator on ants is the larval antlion
(Myrmeleontidae spp.), which captures its prey by
constructing and hiding inside conical traps in the
ground (Coelho 2001). Though ants are extremely
abundant in acacia groves, antlion traps are rarely
found there (Farji-Brener et al. 2008). This
suggests antlion distribution is not primarily
driven by prey availability.
As antlions are dependent on constructed traps
for prey capture, the substrate in which the trap is
built may affect capture success. Due to the high
energetic cost of constructing traps in hard soils
6
(Fertin 2006), antlions build smaller traps there
(Farji-Brener 2003). However, in a finer substrate
such as sand, antlions can build larger and more
efficient traps (Fertin and Casas 2006). Greater
capture success in sandy substrates suggests that
substrate type may have an effect on antlion
distribution.
We tested the hypothesis that substrate type
would be more important than prey availability in
driving antlion distribution. We predicted that
antlions would prefer sandy substrates for trap
construction and that larger traps would increase
prey capture success.
METHODS
Antlion Trap Census
We studied two 60 meter transects in Palo Verde
National Park: one along a sandy roadside and one
through a stand of V. collinsii. We counted the
number of antlion traps in 10 randomly selected 1meter square quadrats for both transects.
Capture Dynamics
To test capture success and efficiency of antlions,
we collected ten P. spinicola and ten ants of other
species. P. spinicola were collected from V.
collinsii branches, and the other species were
collected from fallen logs and tree snags. We
categorized each ant by size class (small < 0.5 cm,
medium 5-1.0 cm, large >1.0 cm). We then
Dartmouth Studies in Tropical Ecology 2014
selected 20 antlion traps, measured their diameter,
and randomly dropped either a P. spinicola or
other species of ant into the trap. We recorded
capture success (whether or not an antlion killed
its prey) and latency to attack.
Substrate Preference
We collected 16 antlions from roadside traps. We
collected sandy substrate from roadsides and hardpacked soil from V. collinsii groves, which we
ground into finer particles. To mimic the packed
soil found under V. collinsii, we created a
moldable mix of soil and water, which we then
dried in the sun. Some of the dried soil was then
reground and sieved to a consistency similar to the
sand. In all cases, we added equal amounts of
water to each substrate to avoid confounding the
treatments. We then evenly divided 16 containers
between pairs of substrates. Eight arenas provided
the choice between sand and hard packed soil,
while the other eight contained sand and ground
soil. We placed one antlion in each arena on the
border between the two substrates and allowed the
antlions 24 hours to build traps. We recorded the
substrate chosen and the trap diameter of all
constructed traps.
Statistical Analyses
We used a two-tailed t-test to determine whether
antlion trap densities differed across substrates in
our census. We used a nominal logistic regression
to assess the antlions’ preference for sand versus
soil substrates. A second nominal logistic
regression was used to analyze the contribution of
ant size and ant species to antlion feeding success.
Finally, we used a nominal logistic regression to
predict antlion feeding success based on trap
diameter. We used JMP 11.0 (SAS Institute, Inc.
2013) for all statistical analyses.
RESULTS
Substrate Preference
Antlions constructed traps in higher densities in
sandy roadsides (mean ± SE = 1.90 ± 0.89 m-2)
than in V. collinsii groves (mean ± SE = 0 ± 0 m-2)
(F1,18= 4.58, P= 0.046; Figure 1). In preference
experiments, antlions preferred sandy substrates
(chi-square= 4.92, P= 0.026), but this preference
was not significantly affected by the choices
provided (chi-square= 4.31, P= 0.34, df= 2, Figure
2).
Capture Dynamics
Antlion capture success was not influenced by ant
type (chi-square = 0; P = 1, df = 1), but was
influenced by ant size, with larger ants more likely
to escape antlion traps than smaller ants (chisquare = 12.98; P = 0.0015, df = 2; Figure 3). Ants
placed in larger traps were less likely to escape
(chi-square = 4.31; P = 0.038, df = 1; Figure 4),
indicating that larger traps led to more successful
feedings.
Figure 1. Antlion trap density was significantly higher in
roadside sandy sites than hard-packed acacia soil sites
(F1,18 = 4.58, P = 0.046). Error bars represent standard
error.
Figure 2. Antlions prefer sand over any type of soil (chisquare = 4.92, P = 0.026, df = 1).
Figure 3. Larger ants (>1.0 cm) were more successful at
escaping antlion traps than small (<0.5 cm) and medium
(0.5-1.0 cm) ants (chi-square = 12.98; P = 0.0015, df =
2). Error bars represent standard error.
7
Palo Verde
Figure 4. Antlions in larger traps had higher success
rates in killing ants. Heavy line represents median, box
represents upper and lower quartiles, and whiskers
represent range.
DISCUSSION
We investigated the importance of substrate type
to antlion distribution. We found a significant
preference for sandy substrate in antlion trap sites
(Figs. 1 and 2). Previous research has shown that
traps built in sand are bigger (Fertin and Casas
2006). We found that larger traps led to more
successful feedings (Figure 4). Therefore,
construction in sandy soils leads to traps with
greater prey capture success. This explains the
greater density of antlion traps on roads than in V.
LITERATURE CITED
Burgess, M.G. 2009. Sub-optimal pit
construction in predatory antlion larvae
(Myrmeleon sp.) Journal of Theoretical
Biology 260: 379-385.
Coelho, J.R. 2001. The natural history and
ecology of antlions (Neuroptera:
Myrmeleontidae). Ex Scientia 7: 3-12.
Crowley, P.H. and M.C. Linton. 1999.
Antlion foraging: Tracking prey across space
and time. Ecology 80: 2271-2282.
Farji-Brener, A.G., D. Carvajal, M.G. Gei, J.
Olano, J.D. Sanchez. 2008. Direct and
indirect effects of soil structure on the density
of an antlion larva in a tropical dry forest.
Ecological Entomology 33: 183-188.
8
collinsii groves, despite the lower abundance of
ants on roads (Crowley 1999). We observed no
difference in antlion capture success between prey
species, confirming that antlion success is not
related to prey type. Therefore, our findings on
substrate preference and capture success are
consistent with our hypothesis that substrate type
is more important than prey availability in driving
antlion distribution.
Many predators are limited in their
distribution and abundance by the quantity and
variety of prey resources. However, when a
predation strategy requires a specific
environmental resource, this may force predator
populations to forgo broader distribution
opportunities.
ACKNOWLEDGEMENTS
We would like to thank R. Calsbeek for the initial
experimental design, V. Venkataraman and J.
Trout-Haney for their guidance and feedback, the
students of the Dartmouth FSP 2014 for their
assistance with the peer review process, and the
staff of Palo Verde National Park for their support.
AUTHOR CONTRIBUTIONS
All authors contributed equally.
Fertin, A. and J. Casas. 2006. Efficiency of antlion
trap construction. The Journal of Experimental
Biology 209: 3510-3515.
Gatti, M.G. and A. G. Farji-Brener. 2002.
Low density of ant lion larva (Myrmeleon
crudelis) in ant-Acacia clearings: high
predation risk or inadequate substrate?
Biotropical 34: 458-462.
Janzen, D.H. 1966. Coevolution of Mutualism
Between Ants and Acacias in Central
America. Evolution 20: 246-275.
Seigler, D.S. and J.E. Ebinger. 2005. New
combinations in the genus Vanchellia
(Fabaceae: Mimosoideae) from the new
world. Phytologia 87: 139-178.
Dartmouth Studies in Tropical Ecology 2014
EXPLAINING THE DISTRIBUTION OF WETLAND MACROINVERTEBRATES:
CONTRIBUTIONS OF HABITAT, PREY AVAILABILITY, AND FEEDING DYNAMICS
ALLEGRA M. CONDIOTTE, NATHANAEL J. FRIDAY, AARON J. MONDSHINE, ADAM N. SCHNEIDER,
AND ZACHARY T. WOOD
Project Designer: Jessica V. Trout-Haney
Faculty Editor: Ryan G. Calsbeek
Abstract: In wetlands, macrophyte beds create a dense structure that can harbor invertebrates from predation and
ultraviolet radiation. Predation pressures on invertebrates can influence their size distributions and cause mass
migrations into and out of shelter. Using a tropical marsh system in Palo Verde, Costa Rica, we tested the hypothesis
that invertebrates would be larger and more abundant in macrophyte-covered areas of the marsh due to a refuge
effect. We also assessed whether invertebrate prey underwent horizontal diel migrations to avoid predators. We
surveyed invertebrates in three microhabitats (tall macrophytes, short macrophytes, and open water) during morning,
afternoon, and night. Invertebrates were more abundant under macrophyte cover, and the invertebrate predator-toherbivore ratio was greatest in the afternoon. These results suggest that macrophytes provide a refuge for
invertebrates, and that larger invertebrate predators are most active at midday. The higher prevalence of
invertebrates under macrophyte cover emphasizes the importance of aquatic plant populations to marsh community
functions.
Key words: aquatic invertebrates, predator-herbivore interactions, horizontal migrations, macrophytes
INTRODUCTION
Trophic interactions in aquatic systems can alter
community dynamics and organism morphology
(Skoglund et al. 2013). When predators select prey
by size they can cause a shift in prey size
distributions, especially when predators are much
larger than their prey. When fish consume
zooplankton, they typically consume the largest
and most conspicuous individuals first.
Diel migrations can provide zooplankton with
a spatio-temporal refuge from predation. In
shallow wetlands where vertical migration is
impossible, zooplankton often exhibit diel
horizontal migration (DHMs). When predators
(e.g., sunfish) rely on visual cues, most
zooplankton are found in open water during the
night and under macrophyte cover or lower in the
water column during the day (Pasternak et al.
2006).
Zooplankton migrate in the opposite direction
in the presence of nocturnal predators such as the
phantom midge chaoborus (Minto et al. 2010).
The spatial heterogeneity of herbivores caused by
migrations and refugia means that grazing may be
limited to specific areas at different times of day.
We tested the hypothesis that invertebrates
would be more abundant and larger in
macrophytic refugia. We also predicted that
invertebrates would undergo horizontal migrations
based on predator presence, with zooplankton
migrating from the tall macrophytes during the
day into the open at night.
METHODS
On 11 January 2014 we surveyed a 24-meter
transect crossing three microhabitats (tall
macrophytes, short macrophytes, and open water)
in the marsh in Palo Verde National Park, Costa
Rica. We took three integrated water column
zooplankton tows (vertical net sweeps through the
entire water column) per microhabitat. Along the
transect, the first eight meters were dominated by
Thalia geniculata (height > 2 meters), followed by
a second eight-meter microhabitat comprised of
short macrophytes, primarily Eichornia crassipes
and Salvinia auriculata. The last eight meters of
each transect were in open water.
We collected samples at three times during the
day: at 8:30 AM, at 4:00 PM, and at 8:30 PM. We
rinsed the samples with ethanol and used a
dissecting microscope to count, identify, and
measure all invertebrates found in the samples. We
identified the invertebrates taxonomically to
family, and classified them as either a predator or
a grazer accordingly.
9
Palo Verde
Statistical Analysis
We used analyses of variance (ANOVA) in R
statistical analysis software to test for effects of
macrophyte cover and time of day on invertebrate
size and density. For all ANOVAs we used
Tukey’s Honestly Significant Differences (HSD)
test to analyze differences between groups post
hoc. We used JMP11 software to conduct
contingency analyses on predator/herbivore and
plankton/non-plankton groups between
microhabitats and collection times.
RESULTS
Invertebrates were 7.2 times more abundant under
macrophytes than in open areas (Figure 1) (F2,24 =
5.89, P = 0.0083). Invertebrate densities in tall and
short macrophyte areas were not significantly
different (Tukeys HSD, P = 1.0). The length of
aquatic invertebrates between tall and short
macrophytes did not differ (Fig. 2, t90 = -0.087, P =
0.93).
The predator/herbivore ratio in all areas was
highest in the afternoon (Fig. 2, 2 = 8.30, df = 2,
P = 0.016). However, neither aquatic invertebrate
abundance nor length changed with time of
collection (F2,24 = 0.54, P = 0.59 and F2,94 = 0.76, P
= 0.47 respectively; Fig. 3).
Figure 2. Predator/grazer composition differed
with respect to collection time (2 = 8.30, P =
0.016, df = 2). The predator:grazer ratio increased
in afternoon samples.Figure 1. There was a nonsignificant relationship between reward level and
ant abundance (ANOVA, F2,8 = 1.93, P = 0.23).
Ant abundance was calculated as the number of
ants on the specified leaf an hour after treatment.
DISCUSSION
Figure 3. Mean length (cm) for aquatic invertebrates
did not change significantly with collection time for
tall and short macrophyte cover (F2,94 = 0.76, P =
0.47).
Figure 1. Invertebrate abundance in macrophytecovered microhabitats was higher than in open
water (F2,24 = 5.89, P = 0.0083). Invertebrate
abundance between tall and short macrophyte
cover was not significantly different (Tukeys HSD,
P = 1.0).
10
Aquatic invertebrates were more abundant in
macrophyte beds than in open water. This might
be the consequence of increased protection in the
structurally complex plant habitat. Time of day did
not influence overall aquatic invertebrate
abundance. However, the relative abundance of
Dartmouth Studies in Tropical Ecology 2014
predators increased significantly during the
afternoon.
At midday, herbivores could be more
susceptible to UV radiation than predators, which
are larger and have a lower surface area to volume
ratio than the herbivores. Therefore, herbivores
may seek shelter and leave the water column while
predators remain in the water column. Fish, which
typically eat invertebrate predators, may also be
less active during midday, which could then allow
invertebrate predators to move freely in the water
column.
Higher densities of invertebrates in
macrophyte beds indicate that macrophyte
structures may influence essential community
processes. For example, invertebrate herbivores
regulate phytoplankton abundance, thereby
filtering the marsh water. According to our results,
most of the prey for fish and other predators live
under the cover of macrophytes, demonstrating
that macrophytes shelter a major food reserve for
predators at higher trophic levels and constitute an
important trophic link between producers and top
consumers.
LITERATURE CITED
Boeing, W.J., D.M. Leech, C.E. Williamson, S.
Cooke, and L. Torres. 2004. Damaging UV
radiation and invertebrate predation:
conflicting selective pressures for zooplankton
vertical distribution in the water column of
low DOC lakes. Oecologia 138: 603-612.
Bronmark, C. and L.A. Hansson. 2005. The
Biology of Lakes and Ponds. Oxford
University Press, Oxford, Britain.
Pasternak, A.F., V.N. Mikheev, and J.
Wanzenbock. 2006. How plankton copepods
avoid fish predation: from individual
responses to variations of the life cycle.
Journal of Ichthyology 46: 220-226.
Skoglund, S., R. Knudsen, and P.A. Amundsen.
2013. Selective predation on zooplankton by
pelagic Arctic charr, Salvelinus alpinus, in six
subarctic lakes. Journal of Ichthyology 53:
849-855.
Minto, W.J., M.S. Arcifa, and A. Perticarrari.
2010. Experiments on the influence of
Chaoborus brasiliensis Theobald, 1901 (Diptera:
Chaoboridae) on the diel vertical migration of
microcrustaceans from Lake Monte Alegre,
Brazil. Brazilian Journal of Biology 70: 25-35.
ACKNOWLEDGEMENTS
We would like to thank Jessica Trout-Haney for
her insight on our project. We would also like to
thank the 2014 Dartmouth Biology FSP group for
peer reviewing our paper. Additionally, we would
like to thank Sergio for reassuring us that we
would not be viciously mauled by voracious
crocodiles. We would like to acknowledge Ryan
Calsbeek for his melodic tunes, which assisted us
in maintaining a stress-reduced environment
conducive to thinking and reflecting on this paper.
AUTHOR CONTRIBUTIONS
All authors contributed equally in the experiment
and writing process.
11
Palo Verde
BEHAVIORAL THERMOREGULATION BASED ON COLORATION IN TROPICAL
BUTTERFLIES
MAYA H. DEGROOTE, ALEXA E. KRUPP, REBECCA C. NOVELLO, CLAIRE B. PENDERGRAST,
AND AMELIA L. RITGER
Project Designer: Ryan Calsbeek
Faculty Editor: Ryan Calsbeek
Abstract: Ectotherms acquire heat from environmental sources rather than from internal metabolic processes.
Variable ambient temperatures require ectotherms such as butterflies to use behavioral mechanisms to regulate their
body temperatures. In the tropical dry forest, sun and shade habitats are a major source of thermal variation. We
predicted that light and dark butterflies would differ in their distribution across light conditions and use of postural
mechanisms because of their varying heat absorption capacities. To test this hypothesis, we performed experimental
and observational studies of body temperature and behavioral differences between light and dark butterflies. Our
results showed that dark butterflies spent less time in the sun and closed their wings more often than light butterflies,
suggesting that dark phenotypes depend more strongly on behavioral cooling mechanisms. Such differences may
have important implications for the foraging capacities of butterflies as well as other ectotherms.
Key words: Behavior, Posture, Thermoregulation, Tropical butterflies
INTRODUCTION
Ectotherms acquire heat from their environment
and are therefore highly sensitive to variation in
ambient temperature (Van Dyck 1997). Butterflies
are ectotherms found across a large latitudinal
gradient, which requires that they tolerate a broad
range of thermal conditions. To maximize time
available for foraging and mating, butterflies
employ behavioral strategies to maintain a body
temperature within the relatively narrow range
(33-38˚C) necessary for flight (Huey and
Kingsolver 1989). Butterfly wings absorb solar
irradiance and transfer heat to the thorax through
small veins. Two thermoregulatory behaviors
commonly observed in butterflies are movement
between shade and sun and wing position
adjustment – opening wings to maximize
absorbance of solar energy or closing wings to
minimize absorbance and facilitate cooling
(Heinrich 1972).
Tropical butterflies have developed a vast array
of wing shapes and colorations to serve diverse
purposes such as camouflage, mate attraction, and
aposematic display (Basset et al. 2011). Despite
such variation in morphology, optimal butterfly
body temperature is consistent across species
(Heinrich 1972). This pattern suggests that
butterflies may achieve stable body temperatures
by spatially distributing across various light
conditions (e.g., sun versus shade) and changing
thermoregulatory behaviors as a function of body
coloration (Clench 1966).
12
We test the hypothesis that light and dark
butterflies behaviorally thermoregulate in different
ways. Specifically, based on the differing heat
absorption capacities of light and dark structural
colors, we predicted that light butterflies would
spend relatively more time in the sun, while dark
butterflies would spend relatively more time in the
shade (Ball 2012). Moreover, we predicted that
light butterflies would spend relatively more time
with their wings open, while dark butterflies
would spend more time with their wings closed.
METHODS
We conducted all research between 08:00 and
12:00 on January 11 and 12, 2014 in the tropical
dry forest of Palo Verde National Park, Costa
Rica. We classified Palo Verde butterfly species as
light or dark based on the dominant shade of their
wing pattern.
Body temperature experiment
We collected four light and four dark butterflies,
asphyxiated the butterflies with acetone, and
mounted them on a Styrofoam board using
dissection pins. We placed the mounted butterflies
in direct sunlight for eight minutes and recorded
their temperatures with a Raytek Raynger MX
temperature gun at two-minute intervals. We then
placed them in the shade for eight minutes,
repeating temperature measurements. We used
these data to verify our underlying assumption that
butterflies have higher body temperatures in the
Dartmouth Studies in Tropical Ecology 2014
sun than in the shade irrespective of color. Given
that our dataset was made up of multiple estimates
of body temperature over time, we began our
analyses using Repeated Measures ANOVA with
each measure of body temperature as a set of
dependent variables, and butterfly coloration and
light treatment as independent variables. However,
because there was no significant effect of the time
component in the RMANOVA, we simplified all
analyses and calculated a mean value for body
temperature, then used this as the dependent
variable in a two way ANOVA. We included
butterfly coloration and light treatment to test for
the effect of each variable on body temperature,
and we tested for a butterfly-coloration  light
treatment interaction.
Body temperature observational study
We measured the body temperatures of stationary
butterflies found in either the sun (n = 17) or shade
(n = 26) using the Raytek temperature gun. We
recorded and averaged three ambient temperature
readings less than 25 cm from the individual. We
then compared body temperature patterns between
our observational and experimental studies.
Spatial thermoregulation observational study
To determine whether butterflies were spatially
distributed based on color, we observed the total
number of light and dark butterflies in sun (n =
155) and shade (n = 95) on both days of sampling.
Observers spent equal time in an open field and
under a forested canopy, recording the color and
light condition of all butterflies seen. We
compared the proportions of time light and dark
butterflies spent in each light condition using a
chi-square test.
Postural thermoregulation observational study
To determine whether butterflies use changes in
body posture to thermoregulate, we measured
wing position of stationary butterflies and
recorded ambient temperature with an Onset Hobo
Pendant temperature logger between 08:00 and
12:00 on January 12, 2014. We defined wing
positioning as either open, with wings touching
dorsally, or closed, with wings outstretched
laterally. We compared wing positions of light
and dark butterflies throughout the day using a
chi-square test. All analyses were performed using
Microsoft® Excel and JMP® 10 statistical
software.
RESULTS
Body temperature experiment
All butterflies, regardless of wing color, were
significantly warmer in the sun than in the shade
(mean Tb (sun) = 48.2C (1.46) vs. mean Tb
(shade) = 32.85C (0.16); ANOVA, F1,13 = 186.63,
p < 0.0001; Fig. 1A). The two factor ANOVA
revealed a significant interaction between butterfly
coloration and light treatment (t3,12 = -2.52,
Figure 1. (A) There was a significant interaction between butterfly coloration and light treatment (t3,12 = -2.52, P
= 0.02) indicating that although there was no difference in butterfly temperature in the shade (32.94°C ±0.95)
dark versus 32.76°C ± 0.95 light), dark butterflies had warmer body temperatures in the sun (50.7°C ±0.95)
relative to light butterflies (45.75°C ± 0.95). (B) Field observations showed that body temperature did not differ
between light and dark butterflies regardless of light condition (ANOVA, F 3,44 = 1.27, P = 0.30). Dashed lines
indicate the optimal range of body temperatures for butterfly flight (33-38˚C).
13
Palo Verde
P=0.02) indicating that although there was no
difference in butterfly temperature in the shade
(32.94°C ±0.95 dark versus 32.76°C ±0.95 light),
in the sun, dark butterflies had warmer body
temperatures (50.7°C ±0.95) compared to light
butterflies (45.75°C ±0.95).
Body temperatures observational study
Field observations showed that body temperature
of butterflies did not differ in terms of either
coloration or light condition (ANOVA, F3,44=1.27,
P=0.30, Fig. 1B).
Spatial thermoregulation observation study
Light butterflies were more commonly found in
the sun, and dark butterflies were more commonly
found in the shade (2=29.071, df=1, P<0.0001,
Fig. 2).
Figure 2. Light butterflies were significantly more
likely to be found in sunny areas, and dark
butterflies were more likely to be found in shady
areas (2=29.07, df=1, P<0.0001, df=1).
Postural thermoregulation observational study
Wing posture differed significantly between
stationary dark and light butterflies (P=0.002, Fig.
3) but did not vary by temperature (P=0.66) or
time (P=0.28). All stationary dark butterflies had
closed wings, whereas light butterflies varied in
their wing posture (Fig. 3).
DISCUSSION
We have shown that the thermal environment
predicts the spatial distribution and postural
behaviors of butterflies based on wing color. We
observed a difference between butterfly
14
Figure 3. Wing posture was significantly different
between stationary dark and light butterflies
(2=9.92, P=0.0016, df=1). Dark butterflies were
never observed with open wings.
temperatures in the sun and shade treatments of
the experiment (Fig. 1A) but not in our field
observations (Fig. 1B). This suggests that butterfly
behavior influences their body temperature. In the
field, butterfly body temperature remained near
the ideal metabolic range (Fig. 1B), while in the
experiment, butterflies in the sun reached lethal
temperatures (Fig. 1A).
We also found that light butterflies were
significantly more likely than dark butterflies to be
found in the sun (Fig. 2). This suggests that
butterfly coloration influences spatial distribution.
Wing posture also differed between dark and light
butterflies, supporting the hypothesis that heat
absorption capacity motivates posturing behavior.
Our findings suggest that dark
tropical butterflies have a physiological
disadvantage in the sun due to their susceptibility
to overheating. This cost is mitigated by
behavioral strategies such as spending more time
in the shade. Although dark coloration is costly in
terms of thermoregulation, it may be advantageous
in terms of camouflage (Sun et al 2013).
Tradeoffs in coloration and thermoregulatory
behavioral strategies may also affect foraging
capacity, a major determinant of butterfly fitness.
By extending the number of hours during which
butterflies can forage, thermoregulatory strategies
offer significant fitness gains (Clench 1966).
Morphological adaptations and behavioral
thermoregulation allow butterflies and other
ectotherms to tolerate a diversity of habitats and
environmental conditions.
Dartmouth Studies in Tropical Ecology 2014
ACKNOWLEDGEMENTS
We would like to thank Jessica Trout-Haney and
Ryan Calsbeek for their help on statistical methods
and insight into the delicate world of the butterfly.
We would also like to thank the 2014 Dartmouth
Biology FSP group or peer reviewing our paper.
We would like to acknowledge the butterflies of
Palo Verde National Station for being very
understanding of the sacrifice of their friends in
the name of science.
LITERATURE CITED
Basset, Y., R. Eastwood, S. Legi, D. Lohman, V.
Novotny, T. Treuer, S. Miller, G. Weiblen, N.
Pierce, S. Bunyavejchewin, W. Sakchoowong,
P. Kongnoo and M. Osorio-Arenas. 2011.
Comparison of rainforest butterfly
assemblages across three biogeographical
regions using standardized protocols. Journal
of Research on the Lepidoptera 44: 17-28.
Ball, Philip (May 2012). Nature's Color
Tricks. Scientific American 306: 74–79.
Clench, H.K. 1966. Behavioral Thermoregulation
in Butterflies. Ecology 46: 1021-1034.
Heinrich, B. 1972. Thoracic Temperatures of
Butterflies in the Field Near the Equator.
Comparative Biochemical Physiology 43: 459467.
Kingsolver, J.G. 1985. Butterfly
Thermoregulation: Organismic Mechanisms
and Population Consequences. Journal of
Research on the Lepidoptera 24: 1-20.
Schowalter, T.D. 2009. Insect Ecology: An
Ecosystem Approach. Elselvier, Burlington,
MA.
Sun, J.Y., B. Bhushan, J. Tong. 2013. Structural
coloration in nature. RSC Advances 35: 1486214889.
Wasserthal, L.T. 1975. The Role of Butterfly
Wings in the Regulation of Body
Temperature. Journal of Insect Physiology 21:
1921-1930.
AUTHOR CONTRIBUTIONS
All authors contributed equally.
15
Palo Verde
FACTORS AFFECTING VIGILANCE BEHAVIOR IN THE PRIMATES
ALOUATTA PALLIATA AND CEBUS CAPUCINUS
ANN M. FAGAN, EMILY R. GOODWIN, LARS-OLAF HÖGER, AND ELLEN R. PLANE
Project Design: Vivek V. Venkataraman
Faculty editor: Ryan G. Calsbeek
Abstract: Animals use vigilance behaviors to detect potential threats and monitor social dynamics. According to the
energy input-vigilance tradeoff hypothesis, energy intake is negatively correlated with vigilance because vigilance
and feeding cannot be performed simultaneously. In larger groups, shared responsibility for predator detection tends
to reduce per capita vigilance. However, it is unclear whether these trends should hold for primates because of their
upright feeding behavior, manual dexterity, and complex social interactions. We monitored feeding and vigilance
activity of Alouatta palliata and Cebus capucinus in Palo Verde National Park, Costa Rica. We tested the hypothesis
that greater manual dexterity in C. capucinus would allow for more compatibility in vigilance and feeding than in A.
palliata. Additionally, we predicted that social monitoring within primate groups would be a stronger driver of
vigilance than predation risk. We found no evidence for an energy intake-vigilance tradeoff in either capuchin or
howler monkeys. Individual vigilance increased with group size in capuchins, but not in howlers, suggesting that
social monitoring is more important for capuchins. Our results suggest that feeding adaptations and social
organization influence the allocation of vigilance in primates.
Key words: Alouatta palliata, Cebus capucinus, energy input-vigilance tradeoff, group size effect on vigilance
INTRODUCTION
Vigilance behaviors allow animals to gauge
predation risk and assess social dynamics.
Vigilance, however, conflicts with other essential
behaviors such as eating and resting; this tradeoff
between energy maximization and vigilance is
documented in multiple taxa (Houston et al. 1993).
Group-living animals share responsibility for
predator detection to reduce per capita energy
spent on vigilance. In many species of birds and
mammals, per capita vigilance decreases with
increasing group size, a phenomenon known as the
group size effect (Elgar 1989). However, in
primates, evidence for both the energy inputvigilance (EI-V) tradeoff and the group size effect
on vigilance is ambiguous. One explanation is that
primates have the ability to feed upright and
collect and process food with their hands, allowing
them to simultaneously feed and monitor their
surroundings (Treves 2000). The absence of the
group size effect on vigilance in some free-ranging
primates (Treves 2000), however, may be due to
the importance of vigilance in social monitoring,
or looking around at other members of the group
(Hirsch 2002).
Free-ranging groups of both white-faced
capuchins Cebus capucinus and mantled howlers
Alouatta palliata inhabit the tropical dry forest of
16
Palo Verde National Park, Costa Rica. As highly
dexterous frugivores/insectivores, capuchins feed
primarily by picking up and holding food items
(Tomblin & Cranford 1994). Folivorous howler
monkeys are less dexterous, pulling branches to
themselves and eating directly from the branch
(Tomblin & Cranford 1994). Increased manual
dexterity in capuchins may enhance their ability to
simultaneously feed and observe their
surroundings.
Because of this difference in feeding modes,
we investigated the hypothesis that the EI-V
tradeoff would be less pronounced in capuchins
than in howlers. Additionally, we expected that
increased conspecific vigilance through social
monitoring among larger groups of primates
would result in increased vigilance with group
size.
METHODS
We monitored A. palliata and C. capucinus
feeding activity in Palo Verde National Park,
Costa Rica over a three day period from 10-13
January, 2014 during the hours of 06:00-12:00.
Once a group was located, we recorded group size
and, when possible, noted each individual as male,
female, or juvenile (sex indistinguishable).
Dartmouth Studies in Tropical Ecology 2014
We began opportunistic focal sampling (Altmann
1974) of feeding bouts and continued until view of
the monkey’s face became obstructed, feeding
activity terminated, or 180 seconds elapsed. We
recorded “looks,” “scans,” and “bites” taken by
monkeys. We defined both looks and scans as a
change in head position; a “look” was an eye
movement directed upwards or away for less than
two seconds, whereas a “scan” was a sweeping
gaze lasting more than two seconds. We recorded
a “bite” when an individual placed food in its
mouth or brought its mouth to a food source. One
observer monitored behavior and dictated looks,
bites, and scans to a second person, who recorded
them. All behaviors were converted to rates.
We stopped monitoring a group when monkeys
were no longer feeding, a group fled, monkeys
became obscured from view indefinitely, or we
had monitored an individual for three feeding
bouts.
Statistical Analyses
We performed linear regression to test the
relationship between feeding rate (bites per
second) to vigilance (looks plus scans per second).
We also used linear regression to determine
whether group size was correlated with vigilance
behaviors per second. Group size was defined as
the number of adults in each group. All statistical
analyses were performed using JMP 11.0 (SAS
Institute, Cary, NC).
RESULTS
Energy Input - Vigilance Tradeoff
There was no significant relationship between
feeding rates and vigilance rates in either capuchin
groups (slope= 0.142 ± 0.06, P = 0.37, r2 = 0.04;
Fig. 1) or howler groups (slope = 0.097 ± 0.03, P =
0.50, r2 = 0.02; Fig. 1).
Group Size Effect on Vigilance
As group size increased, capuchins showed higher
rates of vigilance (slope = 0.016 ± 0.06, P = 0.05, r
2
= 0.19; Fig. 2), whereas vigilance rates of howler
monkeys did not vary with group size (slope =
0.004 ± 0.02, P = 0.25, r2 = 0.25; Fig. 2).
Figure 1. Energy input was not significantly correlated
with vigilance in either capuchins or howlers.
Figure 2. Vigilance increased significantly with group
size in capuchins but not in howlers.
DISCUSSION
Differences in feeding behavior between
capuchins and howlers do not appear to affect the
EI-V tradeoff (Fig. 1). We found no evidence of
an EI-V tradeoff in either species. This result is
consistent with previous findings that feeding and
vigilance are compatible in primates (Treves
2000). The presence of grasping hands in all
primates, rather than interspecific differences in
feeding behavior, may explain the lack of an EI-V
tradeoff. We observed the importance of grasping
hands in individuals of both species, who appeared
to coordinate looks and scans with bites. In many
other taxa (e.g. ungulates), the predominant use of
the head/neck in food harvesting limits the ability
to perform other behaviors simultaneously.
Capuchin group size was positively correlated
with vigilance, but this relationship did not hold
for howlers. Our data support the theory that in
17
Palo Verde
some group-living animals, vigilance is primarily
conspecific (Hirsch 2002). Howler monkeys
generally live in groups with a single-male harem
structure (Wang and Milton 2003), while
capuchins often form multi-male multi-female
groups (Rose and Fedigan 1995). The presence of
multiple males within a group could explain why
we observed increased vigilance with group size in
capuchins but not in howlers. Our results support
the concept that group composition and social
organization influence vigilance allocation.
This study provides evidence that the grasping
hands of primates allow them to avoid sacrificing
vigilance while feeding. Because the complex
social structures of primate groups require
heightened conspecific vigilance relative to other
taxa, primate multitasking is especially beneficial.
LITERATURE CITED
Altmann, J. 1974. Observational study of
behavior: sampling methods. Behavior
49: 227- 267.
Elgar, M.A. 1989. Predator vigilance and group
size in mammals and birds: a critical
review of the empirical evidence. Biological
Reviews 64: 13-33.
Hirsch, B.T. 2002. Social monitoring and
vigilance behavior in brown capuchin
monkeys (Cebus apella). Behavioral Ecology
and Sociobiology 52: 458-464.
Houston, A.I., McNamara, J.M., and J.M.C.,
Hutchinson. 1993. General results concerning
the trade-off between gaining energy and
avoiding predation.
Philosophical Transactions: Biological
Sciences 341: 375-397.
Rose, L.M., and Fedigan, L.M. 1995. Vigilance
in white-faced capuchins, Cebus
capucinus, in Costa Rica. Animal
Behaviour 45: 63-70.
Tomblin, D.C., and Cranford, J.A. 1994.
Ecological niche differences between Alouatta
palliata and Cebus capucinus comparing
feeding modes, branch use, and diet. Primates
35: 265-274.
Treves, A. 2000. Theory and methods in studies
of vigilance and aggregation. Animal
Behaviour 60: 711-722.
18
ACKNOWLEDGEMENTS
We would like to thank V. Venkataraman for the
project concept, the students of the Dartmouth FSP
2014 for their help during the peer review process,
the staff of Palo Verde National Park for their
support, and the monkeys for cooperating.
AUTHOR CONTRIBUTIONS
All authors contributed equally.
Dartmouth Studies in Tropical Ecology 2014
ONTOGENY AND PREDATORY EFFICIENCY IN
CENTRUROIDES MARGARITATUS
ANN M. FAGAN, EMILY R. GOODWIN, AARON J. MONDSHINE,
ELLEN R. PLANE, AND AMELIA L. RITGER
Faculty editor: R. Calsbeek
Abstract: Ontogenetic changes in morphology and predation strategies can impact how well predators capture their
prey. In some species of scorpions, prey capture strategies change with body size. We hypothesized that scorpion
age would be positively correlated with capture efficiency and negatively correlated with use of stingers, since older
scorpions are more experienced and have larger chelae. Using body size as a proxy for age, we explored changes in
predation strategies and capture efficiency with respect to age in the scorpion Centruroides margaritatus. We found
a positive relationship between age and capture efficiency, although this finding may have been confounded by prey
size. We found no relationship between age and use of stinger. This study reveals the need for further research on
scorpion development and its effect on predation efficiency and strategy.
Key words: ontogenetic variation, capture efficiency, predatory strategy, Centruroides margaritatus
INTRODUCTION
Adult animals are generally more successful
predators than juveniles. This is often attributed to
(1) morphological differences and (2) ontogenetic
changes in prey capture strategy (Ringler 1983).
As animal body size and age increase, foraging
and predation efficiency also increase through
either enhanced searching and hunting efficiency,
or through increased prey handling efficiency
(Mittelbach 1981, Breitwisch 1987, Ehlinger
1990).
Bark scorpions (Centruroides spp.) exhibit
direct development, meaning scorplings have adult
body proportions throughout growth (McReynolds
2012). Therefore, size is an appropriate proxy for
age (Polis and McCormick 1986, McReynolds
2012).
Ontogenetic niche shifts in predatory behavior
have been documented in other scorpions (Casper
1985, McReynolds 2012). Younger scorpions
paralyze prey by stinging while older individuals
are able to restrain their prey with chelae alone
(Casper 1985; Figure 1). When older scorpions
restrict stinger use, they conserve energy that
would have been spent on venom production (Rein
1993). This behavioral change indicates that older
scorpions should be more efficient predators.
We hypothesized that the Central American
bark scorpion C. margaritatus would show
ontogenetic variation in both predatory behavior
and morphology. Specifically, we predicted that
older scorpions would (1) have higher overall prey
capture efficiency because of their larger body size
and (2) crush prey in their chelae instead of using
their stingers to maximize predation success while
minimizing energy input.
METHODS
Scorpion & Prey Collection
We collected 15 scorpions and 15 crickets on the
nights of 13-15 January 2014 in Palo Verde
National Park, Costa Rica by searching roads and
building perimeters after sunset. We placed the
scorpions in individual containers, provided them
with water, and starved them for 14-42 hours. We
kept the prey in a separate container. We weighed
each prey item and each scorpion prior to
conducting our trials.
Predation Trials
We conducted predation trials in a dark area
covered in black fabric to reduce ambient light.
Scorpions and prey items were ranked from
smallest to largest and paired accordingly to
ensure that all scorpions received proportionally
sized prey items. One scorpion container was
placed in the trial area, and the trial began when
the prey item was dropped into the container. A
“strike” was defined as a rapid movement of the
scorpion’s pedipalps towards its prey. “Capture”
was defined as the final strike, after which the prey
did not escape from the scorpion. We also
19
Palo Verde
recorded the total number of strikes, the total
number of stings (which always occurred after the
prey had been captured), and the cumulative time
from the beginning of the trial to the time of prey
capture. We used only red lighting inside the trial
area to avoid confounding effects due to scorpion
phototaxis.
RESULTS
Carapace length and capture efficiency were
positively correlated (slope = 0.029 ± 0.16, P =
0.01, r2 = 0.55, Fig. 2). Carapace length did not
differ between the sexes (t = 0.66, P = 0.52, df =
24). There was a significant positive correlation
between prey mass and capture efficiency (slope =
2.49 ± 0.16, P = 0.01, r2 = 0.56). We found no
relationship between carapace length and residual
capture efficiency controlled for prey mass.
There was no significant difference in
carapace length between individuals that stung
their prey compared to those that did not (t = 0.17,
P = 0.87, df = 6, Fig. 2).
Figure 1. Image of C. margaritatus showing measured
morphological traits.
Morphological Measurements
After the predation trials, we sacrificed all
scorpions by asphyxiating them with acetone. We
then sexed each scorpion and measured carapace
length of each individual (Figure 1).
Statistical Analyses
We used a linear regression to test for a
relationship between carapace length and capture
efficiency. We defined capture efficiency as the
ratio of capture success (a binary variable scored 0
or 1) to the number of strikes. To test for a sex
effect on size, we used a two-tailed t-test on mean
carapace length by sex. We also used a linear
regression to examine the relationship between
prey mass and capture efficiency. We used the
residuals from this test to conduct a second linear
regression comparing carapace length and residual
capture efficiency.
To test our hypothesis regarding age and
stinger use, we only used trials where scorpions
struck at least once, assigning a value of 1 to
scorpions that successfully captured their prey
using a sting and a value of 0 to those that
captured prey with no sting. We then conducted a
two-tailed t-test to compare mean carapace length
between scorpions that stung and scorpions that
did not sting. All statistical analyses were
conducted using JMP 11.0 (SAS Institute, Cary,
NC).
20
Capture Efficiency
1
0.8
0.6
0.4
0.2
0
20
30
Carapace Length (mm)
40
Figure 2. Capture efficiency increased with scorpion
carapace length (slope = 0.029 ± 0.16, P = 0.01, r2 =
0.55). We defined capture efficiency as capture success
divided by the number of attempts (0 = no capture
success, 1 = capture success).
No Sting
Sting
20
25
30
35
Carapace Length (mm)
40
Figure 3. There was no difference in mean carapace
length between scorpions that stung prey and scorpions
that did not sting (t = 0.17, P= 0.87, df = 6).
DISCUSSION
Age was not correlated with increased capture
efficiency or decreased stinger use in C.
margaritatus. There was no relationship between
Dartmouth Studies in Tropical Ecology 2014
scorpion size and capture efficiency when
controlling for prey mass. Additionally, we did not
observe a significant relationship between
predation strategy and scorpion age (Fig. 2). This
result suggests that both young and old scorpions
use their stingers to subdue their prey upon
capture. This finding contradicts previous research
indicating that older scorpions crush prey in their
large chelae and use their stingers less frequently
(Casper 1985) to conserve energetically costly
venom (Nisani et al 2007)
A major caveat of the study is the likely
presence of a confounding effect of prey size on
both hypotheses, making it difficult to determine
the relationship between age, capture efficiency,
and predation strategy. Although we saw a
positive correlation between body size and capture
efficiency, this could be because larger crickets
were easier to catch. In addition, a consistently
high prey to scorpion ratio may have compelled
both young and old scorpions to use their stingers.
Our results do not support the hypothesis that
C. margaritatus exhibits ontogenetic variation in
morphology and predatory behavior. Such speciesspecific ontogenetic complexities require further
exploration.
LITERATURE CITED
Casper, G.S. 1985. Prey capture and stinging
behavior in the Emperor scorpion, Pandinus
imperator (Koch) (Scorpiones, Scorpionidae).
The Journal of Arachnology 13: 277-28.
Ehlinger, T.J. Habitat choice and phenotypelimited feeding efficiency in bluegill:
individual differences and trophic
polymorphism. Ecology 71: 886-896.
McReynolds, C.N. 2012. Ontogenetic shifts
in microhabitat use, foraging and temporal
activity for the striped bark scorpion
Centruroides vittatus (Scorpiones: Buthidae).
Euscorpius 144: 1-19.
Mittelbach, G.G. 1981. Foraging efficiency
and body size: A study of optimal diet and
habitat use by bluegills. Ecology 62: 1370–
1386.
Nisani, Z., Dunbar, S.G., and Hayes, W.K.
2007. Cost of venom regeneration in
Parabuthus transvaalicus (Arachnida:
Buthidae). Comparative Biochemistry and
Physiology Part A: Molecular and Integrative
Physiology 147: 509-513.
Polis, G.A., and McCormick, S.J. Patterns of
resource use and age structure among species
of desert scorpion. Journal of Animal Ecology
55: 59-73.
Rein, J.O. 1993. Sting use in two species of
Parabuthus scorpions (Buthidae). The Journal
of Arachnology 21: 60-63.
Ringler, N.H. 1983. Variation in foraging
tactics in fishes. Pages 159-171 in Predators
and prey in fishes. Dr W. Junk. The
Netherlands.
ACKNOWLEDGEMENTS
We would like to thank Sergio Padilla of the
Organization for Tropical Studies for his advice on
capturing and handling scorpions, the park staff
for assisting us in capturing scorpions and J.
Trout-Haney for her help in conducting predation
trials. We would especially like to acknowledge R.
Calsbeek for his valued advice on acceptable
methods of randomization. Thanks for nothing.
AUTHOR CONTRIBUTIONS
All authors contributed equally.
21
Palo Verde
WETLAND MACROPHYTE DIVERSITY AFFECTS FISH ABUNDANCE, MORPHOLOGY, AND
RICHNESS
ALLEGRA M. CONDIOTTE, MAYA H. DEGROOTE, ALEXA E. KRUPP, ADAM N. SCHNEIDER AND ZACHARY
T. WOOD
Faculty Editor: Ryan Calsbeek
Abstract: Structural complexity and heterogeneity can create diverse ecological niches. Habitat heterogeneity can
explain localized species richness, as it provides a greater variety of niches for species to occupy. We tested whether
edge habitats between macrophyte beds and open water would foster greater fish species richness in a marsh in Palo
Verde National Park, Costa Rica. Fish species richness was highest at the edges of the macrophyte beds where the
two habitats converge. We found the greatest fish abundance outside the macrophyte beds, and the largest fish
within the beds. Our findings suggest that edge habitats are areas of higher habitat heterogeneity that enable
increased species richness.
Keywords: macrophytes, habitat diversity, species richness, tropical freshwater fish
INTRODUCTION
Habitat heterogeneity and structural complexity
play a large role in shaping biological
communities. Heterogeneity is an established
explanation for species richness, as it enables
species to specialize on particular niches
(Rosenzweig 1992).
In aquatic environments, structural complexity
facilitates higher fish species richness, as
demonstrated in temperate marine rocky shore
systems (Eriksson et al. 2006). Marsh plants also
typically foster higher densities of invertebrate
prey and limit predator accessibility (Condiotte et
al. 2014). However, macrophytes can also limit
fish movement and cause diel hypoxia. Emergent
plants limit dissolved oxygen levels by drawing
oxygen from the water column back into the
atmosphere during respiration and decreasing
mixing, thus reducing surface gas exchange (Valk
2012). Such tradeoffs may structure fish
communities in the marsh ecosystem.
Our study addresses how structural complexity
and heterogeneity influence fish biodiversity. We
compared fish species richness, abundance, and
size in a tropical marsh in three different
microhabitats: open water, macrophyte edge, and
complete macrophyte cover. We predicted that the
edge habitat would harbor the highest fish species
richness as a consequence of increased structural
heterogeneity. We also expected fish abundance to
be highest in the edge habitat, and fish size to be
smallest in complete macrophyte cover due to the
difficulty of maneuvering in areas with high
22
structural complexity.
METHODS
On January 14 and 15, 2014, between 8:30 and
10:00, we haphazardly placed six to eight minnow
traps in each of three Palo Verde wetland
microhabitats: areas lacking plants that were
greater than a meter from the nearest macrophyte
beds (open), habitats on the borders of the
macrophytes (edge), and habitats fully enclosed by
macrophytes that were over a meter from the
edges (covered). We recorded water depth and
macrophyte height at each site. Dominant
macrophytes included Salvinia, Thalia geniculata,
Neptunia natans, Cyperaceae Eleocharis,
Ceratophyllum muricatum, Pistia stratiotes, and
Eichhornia spp. We collected traps between 16:00
-17:30 and recorded fish species and size.
Statistical Analyses and Modeling
We used ANCOVA and ANOVA to examine
effects of habitat and species on fish length, and of
habitat type on abundance. We used a post hoc
Tukey's HSD Test to compare ANOVA group
means. All analyses were performed in R
statistical analysis software.
RESULTS
We found four species of fish in edge habitats and
three species of fish in both open water and
macrophyte-covered areas. Banded tetras were
dominant in open and edge habitats, and cichlids
were dominant in the covered areas (Figure 1).
Dartmouth Studies in Tropical Ecology 2014
Fish density was significantly affected by
habitat type (Figure 2, ANOVA, F2,18 = 5.58, P =
0.013). Open and edge habitats had higher
densities than the covered areas (p < 0.03 for both,
Tukey’s HSD). There was no significant
difference in fish density between open and edge
habitats (P = 0.94, Tukey’s HSD test).
Fish were larger in macrophyte beds than in
either the open water or edge area (Figure 3;
ANOVA, F2,88 = 5.39, P = 0.0034). This habitatonly effect became nonsignificant when we
considered the effects of species and the
interaction of species with habitat (Table 1),
indicating that size differences are likely caused
Figure 1. Species richness was greatest in edge
habitats, and fish abundance was greatest in edge
and open habitats. Species codes described in
Table 2. Dashed lines indicate species absence.
by naturally larger fish species residing in
macrophytes (i.e., cichlids, Table 2).
Table 1. ANCOVA results for fish length based on
species and habitat. Most of the variation in length
was explained by species in this model.
Effect
df
Mean sq
F
p
Species
4
6.108
8.467
<0.0001
Habitat
Species *
Habitat
2
0.342
0.475
0.62
3
1.827
2.533
0.063
Residuals
81
0.721
DISCUSSION
As predicted, we found higher fish species
richness at the edge of the macrophyte zone than
in either open or covered areas. It was impossible
to compare fish species per trap across habitat site
due to rarefaction limitations (i.e. some traps had
very few species).
Contrary to our hypothesis, mean fish length
was larger under complete macrophyte cover. This
difference was mainly attributed to a species
effect, as relatively large cichlids were dominant
in the macrophyte beds (Fig. 3, Table 2). Both
cichlid species in the marsh have laterally
compressed bodies, which are optimal for
navigating macrophytic habitats compared to the
other fishes’ cylindrical body forms (Webb 1984).
We caught more fish in the open and edge
habitats than under complete macrophyte cover.
Although macrophyte beds offer increased food
and protection, it appears the costs of daily
hypoxia and high structural complexity in
macrophytes outweigh the benefits for most fishes,
thereby inhibiting their presence (Valk 2012, Fig.
2).
There are tradeoffs for organisms living in any
habitat; by living at the intersection of several
microhabitats, organisms may reap the benefits of
multiple niches, allowing them to minimize
tradeoff costs. Those organisms whose ideal
Table 2. Mean size of fish species caught in minnow traps from open, edge, and covered sites.
Fish
Common name
Mean Length (cm)
Standard Error
Rivulus uroflammeas
K
4.55
1.55
Brachyrhaphis olomina
L
4.42
0.286
Herotilapia multispinosa
C1
4.857
0.637
Astyanax aeneus
T
3.636
0.084
C2
7
---
Parachromis dovii*
Not included in the analysis because we caught only one individual.
23
Palo Verde
Figure 2. Fish density was significantly higher in
open and edge habitats than covered areas (p<0.03
for both, Tukey’s Honestly Significant Differences
test). There was no significant difference between
fish density in open and edge habitat (P=0.94,
Tukey’s HSD test).
Figure 3. Fish were significantly larger in the covered
macrophyte habitat than in either the open water or
edge habitat (Tukey HSD p < 0.003 for differences
between open/edge and covered).
niches exist primarily in either open water or
macrophyte beds may take advantage of edge
habitats, where both conditions are accessible.
We found that habitat heterogeneity,
especially at edge habitats, favors increased
species richness. This conclusion has implications
for many marsh systems, where invasive
macrophytes (e.g. water hyacinth) can reduce
heterogeneity within microhabitats. Habitat
homogenization disrupts species composition,
potentially altering the greater ecosystem
structure.
ACKNOWLEDGEMENTS
We would like to thank Vivek Venkataraman and
Jessica Trout-Haney for their assistance in data
collection, and Jess in particular for swimming
face-down in a crocodile-infested marsh. We
would also like to thank the minnows who donated
their time and energy to our research, and the
brave minnow traps who sacrificed themselves in
the name of science, and which are now
presumably serving as take-out meals for
crocodiles.
LITERATURE CITED
Condiotte, A.M., Schneider, A.N., Friday, N.J.,
Mondshine, A.J., Wood, Z.T. 2014. Wetland
and Macrophyte Diversity Affects Fish
Abundance, Morphology, and Richness.
Dartmouth Studies in Tropical Ecology 2014,
In Press
Eriksson, B.K., Rubach, A., and Hillebrand, H.
2006. Biotic habitat complexity controls
species diversity and nutrient effects on net
biomass production. Ecology 87: 246-254.
Mykra, H., Heino, J., Oksanen, J, and Muotka, T.
2011. The stability-diversity relationship in
stream macroinvertebrates: influences of
sampling effects and habitat complexity.
Freshwater Biology 56:1122-1132.
Valk, A. G. 2012. The Biology of Freshwater
Wetlands. Oxford University Press, Oxford,
Britain.
Rosenzweig, M. L.1992. Species diversity
gradients: we know more and less than we
thought. Journal of Mammology 73: 715-730.
Webb, P.W. 1984. Body form, locomotion, and
foraging in aquatic vertebrates. American
Zoologist 24: 107-120.
24
AUTHOR CONTRIBUTIONS
All authors contributed equally. Adam Schneider
identified all fish by species.
Dartmouth Studies in Tropical Ecology 2014
PATTERNS OF NUTRIENT DISTRIBUTION BETWEEN SOIL AND ABOVEGROUND PLANT
BIOMASS IN TROPICAL FORESTS DO NOT HOLD AT THE UNDERSTORY LEVEL
NATHANAEL J. FRIDAY, LARS-OLAF HÖGER, REBECCA C. NOVELLO, AND CLAIRE B. PENDERGRAST
Faculty Editor: Ryan Calsbeek
Abstract: Most nutrients in a tropical forest system are held in aboveground plant biomass, leaving soils relatively
nutrient poor. We tested for this pattern in understory plants and topsoil, using microarthropod richness and
abundance as indicators of soil quality. We found that microarthropod abundance was positively correlated with soil
moisture levels but not with biomass of understory plants. We discuss our results in terms of the limitations of the
common assumption that ecological patterns established at one scale hold at another.
Key words: microarthropods, nutrient distribution, plant biomass, spatial scale
INTRODUCTION
Tropical forests support a large amount of plant
biomass, and yet tropical soils are remarkably
nutrient-poor (Whitaker 1975). Through
adaptations such as the production of
sclerophyllous leaves, root mats, and high rootshoot ratios, rainforest plants extract nearly all
available nutrients from the soil and trap them in
aboveground biomass, resulting in mineral-poor
and infertile soils (Jordan and Herrera 1981). This
pattern of nutrient distribution has been observed
at the scale of canopy and soil interactions
(Vitousek 1986); however, the dynamics of
nutrient cycling at the smaller scale of topsoil and
understory interactions are less well understood
(Yan 2011).
Extraction of available soil nutrients by
understory plants should leave few nutrients in the
topsoil to support microarthropod life. Soil
microorganisms play a vital role in sustaining and
limiting plant productivity through the processing
and consumption of soil nutrients (Van Straalen
1997). Soil microarthropod diversity and
abundance are correlated with and therefore used
as proxies for soil nutrient content (Stork and
Eggleston 1992). We tested the hypothesis that
microarthropod richness and abundance would be
negatively correlated with understory plant
biomass.
METHODS
We collected soil and plant samples in the tropical
dry forest of Palo Verde National Park on January
14, 2014 between 09:00 and 12:00. We sampled
along a 60 meter transect that ran along a moisture
gradient. At 3 m intervals along the transect, we
randomly determined a distance (between 1 and 10
m) and direction from the transect. We then laid a
30 cm square quadrat at the determined location.
We collected all aboveground plant biomass in
each quadrat. Additionally, we took a total of five
soil cores at haphazard locations within the
quadrat, each 5 cm deep and 2 cm in diameter: one
to measure soil moisture levels and four to
measure microarthropod abundance and richness.
We weighed soil moisture cores and biomass
samples for each site and dried them in a drying
oven at 105˚C for 22 hours, by which time the
largest samples had remained at a constant weight
for three hours. We then recorded dry weight for
each soil sample and calculated moisture content
as the percent water by weight of the original
sample. To extract microarthropods from the soil,
we constructed Berlese funnels from plastic cups
and mesh (Barberena-Arias et al. 2012). We filled
each cup with 30 ml of water and hung the funnels
under 150 watt fluorescent light bulbs for 12
hours. We then identified the microarthropods to
order using a dissection scope.
We performed two multiple regression
analyses using JMP10 analytical software. One
analysis tested for relationships between
aboveground plant biomass and microarthropod
richness and abundance, correcting for soil
moisture. The second analysis tested for
relationships between soil moisture and
microarthropod richness and abundance,
correcting for aboveground plant biomass.
25
Palo Verde
Figure 1. Soil moisture was positively correlated with both (A) microarthropod order richness (multiple
regressions analysis, slope = 48.79  15.95, P = 0.01, r2 = 0.46) and (B) abundance (slope = 55.80 
22.69, P = 0.03, r2 = 0.36), accounting for plant aboveground biomass.
RESULTS
There was no relationship between aboveground
plant biomass and either microarthropod order
richness (Table 1, slope = 0.04  0.09, P = 0.67, r2
= 0.01) or microarthropod abundance (Table 2,
slope = 0.04  0.13, P = 0.77, r2 = 0.01), even after
accounting for soil moisture. However, soil
moisture was positively correlated with
microarthropod order richness (Fig. 1A, Table 1,
slope = 48.79  15.95, P = 0.01, r2 = 0.46) and
abundance (Fig. 1B, Table 2, slope = 55.80 
22.69, P = 0.03, r2 = 0.36) when accounting for
plant biomass.
Table 1. Soil moisture predicted microarthropod order
richness, but aboveground plant biomass did not.
Table 2. Soil moisture predicted microarthropod
abundance, but aboveground plant biomass did not.
26
DISCUSSION
Our results suggest that the patterns of nutrient
storage observed at the broader canopy level of
tropical dry forests do not hold when applied to
understory microhabitats. There was no
relationship between understory plant biomass and
either soil microarthropod abundance or order
richness. Resource limitations, herbivore pressure,
and niche exploitation opportunities vary widely
with ecosystem type and scale. It is therefore
likely that so il-plant nutrient economies within
tropical forests differ depending on spatial scale
and location. These results may also
be explained by variables that we omitted, such as
pore size and soil pH, both of which may influence
microarthropod communities (Hågvar 1990,
Vreeken-Buijs et al. 1998).
We found that soil moisture was positively
correlated with both microarthropod abundance
and order richness. Studies have shown that
fluctuations in soil moisture have considerable
influence on soil fauna (Frouz 2004), which
facilitate mineralization and nutrient cycling (Cole
2004). Our findings suggest that under conditions
of water stress, as found in a tropical dry forest,
moisture is a primary driver of soil nutrient
availability for both plant and microorganism
communities. Future studies should explore in
greater depth the impact of temporal and spatial
variations of moisture on nutrient cycling,
Dartmouth Studies in Tropical Ecology 2014
microarthropod abundance and diversity, and
trophic interactions at the understory level.
ACKNOWLEDGEMENTS
We would like to thank R. Calsbeek, V.
Venkataraman and J. Trout-Haney for their
guidance and feedback, the students of the
LITERATURE CITED
Barberena-Arias, M., G. Gonzalez, and E. Cuevas.
2012. Quantifying Variation of Soil
Arthropods Using Different Sampling
Protocols: Is Diversity Affected? Tropical
Forests, Dr. Padmini Sudarshana (Ed.)
Cole, L., K. Dromph, V. Boaglio, R. Bardgett.
2004. Effect of density and species richness of
soil mesofauna. Biology and Fertility of Soils
39: 337-343.
Frouz, J., A. Ali, J. Frouzova, R. Lobinske. 2004.
Horizontal and vertical distribution of soil
macroarthropods along a spatio-temporal
moisture gradient in subtropical central
Florida. Environmental Entomology 33: 12821295.
Hågvar, S. 1990. Reactions to soil acidification in
microarthropods: Is competition a key factor?
Biology and Fertility of Soils 9:178-181.
Jordan C.F., Herrera R. 1981. Tropical Rain
Forests: Are Nutrients Really Critical? The
American Naturalist 117: 167-180.
Montagnini, F. 2000. Accumulation in aboveground biomass and soil storage of mineral
nutrients in pure and mixed plantations in a
Dartmouth FSP 2014 for their assistance with the
peer review process, and the staff of Palo Verde
National Park for their support.
AUTHOR CONTRIBUTIONS
All authors contributed equally.
humid tropical. Forest Ecology and
Management 134: 257-270.
Stork, N. E. and P. Eggleton. 1992. Invertebrates
as Determinants and Indicators of Soil
Quality. American Journal of Alternative
Agriculture 7: 38-45.
Van Straalen, N.M. 1997. Biological indicators of
soil health. pp. 235-264, Pankhurst, C. Doube,
B. M.: Gupta, V. V. S. R., editors. Vrije
Universiteit, Department of Ecology and
Ecotoxicology, De Boelelaan 1087, 1081 HV
Amsterdam, Netherlands.
Vitousek, P.M. and R.L. Sanford. 1986. Nutrient
Cycling in Moist Tropical Forest. Annual
Review of Ecological Systems 17: 137-67.
Vreeken-Buijs, M.J., Hssink, J., Brussaard L.
1998. Relationships of soil microarthropod
biomass with organic matter and pore size
distribution in soils under different land use.
Soil Biology and Biochemistry 30: 97-106.
Yan, S., A.N. Singh, S. Fu, C. Liao, S. Wang, Y.
Li, Y. Cui, and L. Hu. 2011. Soil fauna index
for assessing soil quality. Soil Biology and
Biochemistry 47: 158-165.
27
Monteverde
ORCHIDACEAE DIVERSITY IN TROPICAL PRIMARY AND SECONDARY FOREST
ANN M. FAGAN, ALEXA E. KRUPP, AARON J. MONDSHINE,
AND AMELIA L. RITGER
Faculty Editor: Ryan Calsbeek
Abstract: Higher biodiversity is consistently found in primary forests, which have more complex structures and
established biological communities. In particular, disturbances associated with secondary forests may have more
acute long-term effects on floral diversity than other taxa. Epiphytic orchids may be used as a measure of floral
diversity because of their influence on ecosystem dynamics and sensitivity to disturbance. We examined orchid
genera composition and diversity in the primary and secondary forests of Monteverde Cloud Forest Biological
Reserve. We found that the secondary forest had higher overall orchid diversity than the primary forest, but that the
diversity per branch by forest type was not significantly different. We suggest that higher orchid diversity is not due
to light gap dynamics but rather other biotic factors.
Keywords: biodiversity, primary forest, secondary forest, Orchidaceae
INTRODUCTION
Structural differences between primary and
secondary forests have important implications for
patterns of biodiversity. New secondary forests
contain light gaps that allow fast-growing
heliophilic plants to flourish (Guariguata and
Ostertag 2001), whereas primary forests contain
complex structures that are able to sustain
established, highly diverse communities (Martin et
al 2004). More tree and liana species are unique to
primary forests than species of all other taxa
(Barlow et al 2007). Thus anthropogenic
disturbances often associated with secondary
forests may have more acute long-term effects on
floral diversity relative to other taxa.
Epiphyte communities are major contributors
to forest biodiversity and influence ecosystem
dynamics by intercepting rainfall, absorbing
nutrients and creating niche habitats (Batke
2012, Shashidhar and Kumar 2009). Epiphytic
species composition differs greatly between
primary and secondary forests (Barthlott et al
2000). In particular, epiphytic orchids can be used
to understand the implications of forest
disturbance because of their dependence on
specific microhabitats and specialized pollinators,
both of which are highly sensitive to disturbance
(Hundera et al 2012). Orchids can also be used as
indicators of biodiversity because of their diverse
physiological adaptations to fit many niches and
because of their intricate relationships with other
plant and animal communities (Shashidhar and
Kumar 2009).
28
To examine the relationship between forest
type and floral diversity, we surveyed epiphytic
orchids in a primary and secondary forest. We
tested the hypothesis that orchid diversity would
differ by forest type, predicting that the primary
forest would contain the greatest orchid diversity.
METHODS
Data Collection
We conducted all research between 0900 and 1700
on 22 and 23 January 2014 in the Monteverde
Cloud Forest Biological Reserve, Costa Rica. Park
scientists showed us the boundaries of the primary
and secondary forests, noting that the secondary
forest was disturbed approximately 30 years ago.
For each forest type, we walked along research
trails and off trail, haphazardly selecting fallen
branches that held epiphytic orchids. Fallen
branches served as proxies for standing trees
because arboreal orchid species grow slowly and
are able to continue growing on a fallen branch
(Zotz 1998).
For each branch, we measured branch length,
circumference, number of orchid species and
number of individuals of each species. We also
took four canopy photographs at each branch site
and recorded tree density, defined as the number
of trees above eye-level within a five-meter radius.
Sample Processing
We used ImageJ (NIH, Bethesda, MD) to analyze
average percent canopy cover at each branch
Dartmouth Studies in Tropical Ecology 2014
Figure 1. There was no significant difference in
orchid genera diversity on branches between
primary and secondary forests (t52=0.16,
P=0.87). Genera diversity was calculated using
the Simpsons diversity index, an index that ranks
diversity from zero to one, zero being most diverse
and one having no diversity.
location. We used both percent canopy cover and
tree density as estimates of light availability. We
also identified orchids to the genus and species
level with the help of a local orchidologist.
Statistical Analysis
We calculated generic diversity within each forest
type using Simpson's diversity index, which
characterizes community biodiversity on a scale
from 0 (infinite diversity) to 1 (no diversity). We
then compared genus composition between the
two forest types using a contingency analysis. We
also conducted an ANOVA to compare canopy
cover and tree density in both forest types. All
analyses were performed using the Al Young
Studios Biodiversity Calculator and JMP® 10
statistical software (SAS Institute, Cary, NC).
Figure 3. Across forest type, there was no
relationship between tree density and number of
orchids per branch (A; slope = 0.02 ± 0.05, P =
0.73, r2 = 0.02) or number of genera per branch
(B; slope = 0.004, P=0.68, r2 = 0.003).
RESULTS
Using the Simpson diversity index, we calculated
a primary forest genera diversity of 0.435 and a
secondary forest genera diversity of
0.291. However, we found no significant
difference in orchid diversity on branches between
Figure 2. There was no relationship between % canopy cover and number of orchids per branch (A; slope = 0.24 ± 0.15, P = 0.11, r2 = 0.0477) or number of genera per branch (B; slope= -0.01 ± 0.03, P = 0.67, r2 =
0.004).
29
Monteverde
Figure 4. Genera composition was different between primary and secondary forests.
the primary (mean ± SE = 0.93 ± 0.05) and
secondary (mean ± SE = 0.94 ± 0.05) forest
(t52=0.16, P=0.87).
We found no significant difference in tree
density between primary forest (mean ± SE =
61.11 ± 3.41 trees) and secondary forest (mean ±
SE = 57.44 ± 3.47 trees; F1,53= 0.567, P =
0.46). We also found no significant difference in
canopy cover between primary forest (mean ± SE
= 20.69 ± 1.06 % canopy cover) and secondary
forest (mean ± SE = 22.59 ±1.10 % canopy cover;
F1,54 = 1.551, P = 0.22). Additionally, we found
that the number of individual orchids per branch
sampled between primary forest (mean ± SE =
4.62 ± 1.18 orchids) and secondary forest (mean ±
SE=6.59 ± 1.22 orchids) was not significantly
different (F1,54 = 1.35, P=0.25).
Canopy cover had no significant effect on
number of orchids (slope = -0.24 ± 0.15, P = 0.11,
r2 = 0.05) or number of genera (slope= -0.01 ±
0.03, P = 0.65, r2 = 0.004) per branch across both
forests (Figure 2A,B). Similarly, tree density had
no significant effect on number of orchids found
30
per branch (slope = 0.02 ± 0.05 P = 0.73, r2 = 0.02)
or number of genera per branch (slope = 0.004,
P=0.68, r2 = 0.003) across forest type (Figure
3A,B). We also found a difference in genera
composition between the primary and secondary
forest (Figure 4).
DISCUSSION
Contrary to our hypothesis, diversity in the
secondary forest was higher than in the primary
forest. The secondary forest of our study area was
logged until approximately 30 years ago.
Depending on how complete this process was, the
disturbance might have enabled a higher level of
orchid diversity in the secondary forest than in the
primary. However, this difference in diversity
between forest types was not represented at the
branch scale. This indicates that orchids are not
evenly dispersed throughout the forest.
The difference in diversity between forest
types may be explained by the intermediate
disturbance hypothesis. This hypothesis posits that
forest diversity may increase following non-
Dartmouth Studies in Tropical Ecology 2014
catastrophic disturbances as new niches from light
gaps are generated for previously excluded
species (Molino and Sabatier 2001). Still, we
found no difference in canopy cover or tree
density between the two forest types, indicating
that light availability may not be the mechanism
driving increased orchid diversity in the secondary
forest. Some structural differences between the
primary and secondary forest of Monteverde may
no longer exist; after multiple decades, secondary
tropical forests can resemble old growth forests in
many ways, including light availability (Martin
2004).
There may still be important biotic differences
between the two forests that impact orchid
community structure. Changes in community
structure may be prompted by pollinator species
composition, seed dispersal or recruitment
limitations (Pauw & Bond 2011, Hubbel et al
1999). These biotic factors may explain the
observed difference in orchid genus composition
between the primary and secondary forest (Figure
4).
Our study adds to the growing debate about
the specific mechanisms driving changes in
biodiversity. Although moderate disturbance may
increase plant diversity in the tropical forest, this
may not be a function of light availability.
Understanding which factors drive biodiversity in
forest ecosystems can inform future land use
practices and conservation strategies.
LITERATURE CITED
Barlow, J., Gardner, T. A., Araujo, I.S., ÁvilaPires, T.C., Bonaldo, A. B., Costa, J.E.,
Esposito, M.C., Ferreira, L.V., Hawes, J.,
Hernandez, M.I.M., Hoogmoed, M.S., Leite,
R.N., Lo-Man-Hung, N.F., Malcom, J.R.,
Martins, M.B., Mestre, L.A.M., MirandaSantos, R., Nunes-Gutjahr, A.L., Overal, W.L,
Parry, L., Peters, S.L., Riveiro-Junior, M.A.,
da Silva, M.N.F., da Silva Motta, C. and C.A.
Peres.2007. Quantifying the biodiversity value
of tropical primary, secondary, and plantation
forests. Proceedings of the National Academy
of Sciences of the United States of America
104: 18555-18560.
Barthlott, W., Schmit-Neuerburg, V., Nieder, J.,
and S. Engwald. 2001. Diversity and
Abundance of Vascular Epiphytes: A
Comparison of Secondary Vegetation and
Primary Montane Rain Forest in the
Venezuelan Andes. Plant Ecology 152: 145156.
Batke, S.P. 2012. A preliminary survey of
epiphytes in some tree canopies in Zambia and
the Democratic Republic of Congo. African
Journal of Ecology 50: 343-354.
Hubbell, S. P., Foster, R.B., O’Brien, S.T., Harms,
K.E., Condit, R., Wechsler, B., Wright, S.J.
and S. Loo de Lao. 1999. Light gap
disturbances, recruitment limitation, and tree
diversity in a Neotropical forest. Science 283:
554-557.
Hundera, K., Aerts, R., Beenhouwer, M.D.,
Overveld, K.V., Helsen, K., Muys, B. and O.
Honnay. 2013. Both forest framentation and
coffee cultivation negatively affect epiphytic
orchid diversity in Ethiopian moist evergreen
afromontane forests. Biological Conversation
159: 285-291.
Hobbs, R.J., and Huenneke, L.F. 1992.
Disturbance, diversity, and invasionimplications for conservations. Conservation
Biology 6: 324-337.
Guariguata, M.R. and R. Ostertag. 2001.
Neotropical secondary forest succession:
changes in structural and functional
characteristics. Forest Ecology and
Management 148: 185-206.
Martin, P.H., Sherman, R.E., and Fahey, T.J.
2004. Forty years of tropical forest recovery
from agriculture: structure and floristics of
secondary and old-growth riprarian forests in
the Dominican Republic. Biotropica 36: 297317.
Molino, J.F. and D. Sabatier. 2001. Tree
diversity in tropical rain forests: A validation
of the intermediate disturbance hypothesis.
Science 294: 1702-1704.
ACKNOWLEDGEMENTS
We are indebted to Gabriel Barboza for his
assistance in locating and identifying orchids and
for his knowledge of the Monteverde cloud forest.
We would also like to thank the Dartmouth FSP
2014 group for their help with the peer-editing
process.
AUTHOR CONTRIBUTIONS
All authors contributed equally.
31
Monteverde
Pauw, A. and W.J. Bond. 2011. Mutualisms
matter: pollination rate limits the distribution
of oil-secreting orchids. Oikos 120: 15311538.
Roberts, M.R. and F.S. Gilliam. 1995. Patterns
and mechanisms of plant diversity in forested
ecosystems: implications for
forest management. Ecological Applications 5:
32
969-977.
Shashidhar, K. S. and A. N. Arun Kumar. 2009.
Effect of Climate Change on Orchids and their
Conservation Strategies. The Indian Forester
135.
Zotz, G. 1998. Demography of the epiphytic
orchid, Dimerandra emarginata. Journal of
Tropical Ecology 14: 725-741
Dartmouth Studies in Tropical Ecology 2014
THERMOREGULATORY COST OF DOMINANCE BEHAVIORS IN HUMMINGBIRDS
ALLEGRA M. CONDIOTTE, EMILY R. GOODWIN, REBECCA C. NOVELLO,
AND ADAM N. SCHNEIDER
Faculty editor: Ryan G. Calsbeek
Abstract: Small endotherms devote a large proportion of their energy to maintaining a constant body temperature. In
hummingbirds, which have the highest per-mass metabolic rate of all vertebrates, feeding competition has a strong
impact on behavior and foraging success and is frequently expressed through energetically expensive dominance
displays. We hypothesized that in a warm tropical climate, dominance behaviors, which produce excess metabolic
heat, would decrease as temperature increased. We tested this hypothesis by observing the behavior of
hummingbirds at a single feeder in the Monteverde Cloud Forest Biological Reserve in Costa Rica. There were
fewer dominance behaviors with increasing temperature, supporting our hypothesis and suggesting that the risk of
overheating outweighs the competitive advantage of frequent dominance behaviors. Even in the tropics, where
temperature ranges are narrow, temperature and other abiotic conditions may create physiological constraints that
drive foraging strategies in small endotherms.
Key words: endothermy, dominance behavior, hummingbirds, metabolism, thermoregulation
INTRODUCTION
Small endotherms must expend a high proportion
of their energy intake in thermoregulation (Gass et
al. 1999). Because of their small body size and
energetically costly form of flight, hummingbirds
have among the highest mass-specific metabolic
rate of all vertebrates (Suarez and Gass 2002).
Energy intake is a limiting factor on hummingbird
reproduction, and fitness is largely dependent on
foraging success (Ydenberg 1998). Foraging
competition has a strong influence on
hummingbird behavior, demonstrated by frequent
inter- and intraspecific displays of dominance
(Stiles and Wolf 1970).
These displays are valuable to ensuring access
to feeding sites (Stiles and Wolf 1970), but they
are also energetically costly and produce excess
metabolic heat that could be detrimental at high
ambient temperatures (González-Gómez et al.
2011). González-Gómez et al. (2011) proposed a
model predicting that hummingbird aggression
level follows a parabolic curve, dropping off at
high and low temperatures (Fig. 1). Since ambient
temperatures in the Monteverde Cloud Forest fall
in the upper temperature ranges of their study, we
hypothesize that the costs of excess metabolic heat
within this range may drive observable patterns of
dominance behavior. We tested the hypothesis that
hummingbird dominance behaviors would
decrease as ambient temperature increased in the
cloud forest.
Figure 1. Thermal performance curve for
hummingbird aggression level adapted from
Gonzalez-Gomez et al. (2011). The curve
represents the theoretical prediction and points
represent empirical data using Sephanoides
sephanoides.
METHODS
We monitored hummingbird activity at a single
hummingbird feeder at Cafe Colibri in the
Monteverde Cloud Forest Biological Reserve,
Costa Rica, from January 21 to 23, 2014.
Hummingbirds were habituated to human presence
in the feeding area, allowing us to observe them
closely without disturbing their activities.
33
Monteverde
Hummingbirds at the feeders included the Brown
Violet-ear (Colibri delphinae), Coppery-headed
Emerald (Elvira cupreiceps), Green-crowned
Brilliant (Heliothryx barroti), Green Violet-ear
(Colibri thalassinus), Magenta-throated Woodstar
(Calliphlox bryantae), Purple-throated Mountain
Gem (Lampornis (castaneoverntris)
cinereicauda), Striped-tailed Hummingbird
(Eupherusa eximia), and Violet Sabrewing
(Campylopterus hemileucurus). Using a HOBO
temperature logger, we recorded ambient
temperature at the birdfeeder for the duration of
the study.
We recorded a three-minute video of
hummingbird activity every half hour between
05:30 and 17:30 each day using a digital video
camera (Nikon D600). We viewed the footage at
25% original speed and counted all hummingbird
visits to the feeder. We categorized all observed
aggressive interactions as dominance displays, and
classified them into three actions: vocalizations
directed towards other birds (chirps), flying
swoops made in the direction of another individual
(chases), and chases ending in collisions
(contacts).
Statistical Analyses
We performed a linear regression between total
weighted dominance behaviors and ambient
temperature. We weighted each of the three
behaviors according to its energetic cost, with
chirps scaled by a factor of 2, chases by a factor of
3, and contacts by a factor of 5 (González-Gómez
et al. 2011). We corrected for number of visits to
the feeder by saving the residuals from a
regression of the number of visits versus weighted
dominance. We performed all statistical analyses
in JMP 10 (SAS Institute).
RESULTS
Time of day and temperature were significantly
correlated (r = 0.69, P < 0.0001, N = 26), and we
chose to use temperature over time of day in our
analysis because of its direct and established
implications for hummingbird thermoregulatory
physiology. Ambient temperature significantly
predicted dominance behaviors, when corrected
for number of visits to the feeder (P = 0.04, r2 =
0.16; Fig. 2).
34
Figure 2. Ambient temperature significantly
predicted weighted dominance behaviors,
corrected for number of visits to feeder (P = 0.04,
r2 = 0.16).
DISCUSSION
Dominance behaviors are an important strategy in
maximizing foraging success by facilitating access
to feeding sites (Stiles and Wolf 1970), but they
may come at a thermoregulatory cost. We found
that ambient temperature significantly predicted
hummingbird dominance (Fig. 2), suggesting that
the risk of overheating may outweigh the benefits
of frequent dominance behaviors at high ambient
temperatures.
Temperatures at Monteverde fall between 13
and 17˚C, representing the upper end of the 520˚C range shown in the model proposed by
González-Gómez et al. (2011, Fig. 1). Therefore,
our results may fit with the trend predicted on the
right side of their parabolic thermal performance
curve.
The patterns of behavioral thermoregulation
shown here may be expected in temperate
endotherms that experience a wide range of
temperatures. However this pattern is less well
understood in areas such as the tropics where
temperature ranges are much narrower. With their
small body mass and high metabolic rate, it
appears that hummingbirds may be forced to
adjust their behavior in response to relatively
small ambient temperature changes to maintain a
stable body temperature. Therefore, the
physiological constraints associated with abiotic
conditions could have important implications for
Dartmouth Studies in Tropical Ecology 2014
foraging strategies in small endotherms across
latitudinal gradients.
for lending us his camera, and Cafe Colibri for
allowing us to conduct research on their property.
ACKNOWLEDGEMENTS
AUTHOR CONTRIBUTIONS
LITERATURE CITED
Evolutionary Ecology 5:220-230.
Stiles, F.G. and Wolf, L.L. 1970. Hummingbird
territoriality at a tropical flowering tree. The
Auk 87: 467-491.
Suarez, R.K. and Gass, C.L. 2002. Hummingbird
foraging behavior and the relation between
bioenergetics and behavior. Comparative
Biochemistry and Physiology 133: 335-343.
Ydenberg, 1998. R.C. Behavioral decisions about
foraging and predator avoidance. In Cognitive
Ecology, ed. R. Dukas, 343-378. Chicago, IL:
Chicago Univ. Press.
We would like to thank R. Calsbeek for his
invaluable advice on our project, V. Venkataraman
Gass, C.L., Romich, M.T., and Suarez, R.K. 1999.
Energetics of hummingbird foraging at low
ambient temperature. Canadian Journal of
Zoology 77: 314-320.
Gonzalez-Gomez, P.L., Ricote-Martinez,
N.,Razeto-Barry, P., Cotoras, I.S., and
Bozinovic, F. 2011. Thermoregulatory cost
affects territorial behavior in hummingbirds: a
model and its application. Behavioral Ecology
and Sociobiology 65: 2141-2148.
Lima, S.L. 1991. Energy, predators and the
behavior of feeding hummingbirds.
All authors contributed equally
35
Palo Verde
OPTIMIZING LEAF INCLINATION ANGLES: WATER SHEDDING AND PHOTOSYNTHESIS IN
CLOUD FOREST UNDERSTORY PLANTS
MAYA H. DEGROOTE, CLAIRE B. PENDERGRAST, AND ELLEN R. PLANE
Faculty Editor: Ryan G. Calsbeek
Abstract: Heterogeneous environments and resource limitation in tropical cloud forests result in highly diverse plant
communities. Plants have evolved different physiological mechanisms to maximize fitness under these conditions. A
plant’s leaf inclination angle affects both its ability to photosynthesize and to shed water, the latter of which is
necessary to prevent epiphyte colonization and facilitate transpiration. Plants with steeper leaf inclination angles
sacrifice photosynthetic capacity to improve water shedding. We predicted that the leaf inclination angle of
understory plants would vary in response to local light conditions. We used experiments and field observations to
investigate the tradeoff between photosynthetic activity and water shedding in the understory plants Piper sp.,
Razicea spicata, and Begonia involucrata. Our results showed that on a species-wide level, understory plants
demonstrate leaf inclination angles suited to the light and moisture conditions of their preferred habitats. This
suggests a role for environmental heterogeneity as a driver of species diversity in tropical forest environments.
Key words: Begonia involucrata, Piper sp., Razisea spicata, leaf inclination angle, canopy cover, water shedding
INTRODUCTION
In the tropical cloud forest, low light availability
inhibits photosynthesis (Rozendaal et al. 2006),
while high rainfall and humidity threaten plant
fitness by limiting transpiration and encouraging
epiphyte growth on leaves (Peabody 1985).
Understory plants maximize photosynthetic
capacity through a variety of physiological
strategies (Rozendaal et al. 2006). However,
prioritizing photosynthesis often requires plants to
sacrifice leaf water shedding capacity. This tradeoff is central to many aspects of leaf morphology,
including angle of inclination (Meng et al. 2014).
Buildup of water on plant leaves reduces plant
fitness. Water film reduces leaf surface
temperature, slowing transpiration and thereby
impairing nutrient transport (Dean and Smith
2014). Standing water on leaves also promotes
growth of epiphytes, which intercept light and
reduce photosynthetic capacity (Lücking and
Bernecker-Lücking 2005). In addition, the added
weight of water and epiphytes requires the plant to
invest more energy in leaf structure (Dean and
Smith 2014).
Tropical plants exhibit adaptations that
increase water shedding, including drip tips, waxy
leaves, and steep leaf inclination angle (Dean and
Smith 2014, Meng et al. 2014). Greater leaf
surface area, height from the ground, and
chlorophyll density increase photosynthetic
capacity, as does a more horizontal leaf inclination
36
angle (Meng et al. 2014). Leaf inclination angle is
influenced by factors such as light availability and
moisture conditions (Lightbody 1985).
Photosynthetic capacity reaches a maximum as
leaf inclination approaches zero, regardless of
plant species (Falster and Westoby 2003).
Optimal leaf inclination angle requires a
compromise between photosynthetic capacity and
water shedding ability (Meng et al. 2014). We
predicted that plants in open canopy environments
would prioritize water shedding through steeper
leaf inclination angles because light is abundant.
We expected plants in denser canopy
environments to have near-horizontal leaf
inclination angles to maximize photosynthesis in
light-limited conditions.
METHODS
Observational study
We selected the tropical cloud forest understory
plants Piper sp., Razicea spicata and Begonia
involucrata as our focal species. Piper sp. and R.
spicata have oval shaped leaves with a single drip
tip, while B. involucrata leaves are asymmetrical
and have two to seven drip tips (Zuchowski 2005).
We haphazardly selected ten plants of each species
in the Monteverde Cloud Forest Biological
Reserve, under a variety of canopy cover
conditions. We photographed the canopy directly
above each plant. We then randomly selected five
leaves on the plant and took a photograph of each
Dartmouth Studies in Tropical Ecology 2014
leaf, holding a plumb line at the point where the
leaf met the petiole to use as a vertical reference.
We also measured the height of each leaf above
the ground.
We used ImageJ software to determine percent
canopy cover and the inclination angle for each
leaf. We determined the angle of inclination with
respect to the horizontal by drawing a line
connecting the leaf’s apex to the point where the
petiole met the leaf, using the plumb line as a
vertical reference.
Experiment
We haphazardly selected two or three undamaged
leaves of each species to determine theoretical
optimum water shedding angles. We positioned
leaves at ten angles between -60° and +60°,
randomizing the order in which these angles were
tested. Before securing the leaf at each angle, we
patted the leaf dry and weighed it. We pipetted a
known mass of water directly onto the leaf from
base to apex, fully coating the leaf’s surface. After
one minute, we weighed the leaf again and used
the difference in mass between dry and wet
measurements, divided by the total water mass, to
determine the ratio of water retained. This was
subtracted from one and converted to a percentage
to determine percent water shed at each angle.
analysis to test for a difference in mean leaf
inclination angle between species.
To interpret the experimental data, we plotted
angle of inclination versus percent water shed and
used polynomial regression to model leaf water
shedding efficiency. Statistical analyses were
conducted using JMP 11.0 (SAS Institute, Cary,
NC) and polynomial fits using Microsoft Excel.
RESULTS
There was no correlation between canopy cover
and leaf inclination angle for any species (Table 1,
Fig. 1). There was also no correlation between leaf
height and leaf inclination angle for any species
(Table 2, Fig. 2).
Mean leaf inclination angle was -18.0° ± 3.9
(± 1SE) for Piper sp., -23.6° ± 3.5 (±1SE) for R.
spicata, and -39.6° ± 2.2 (± 1SE) for B.
involucrata (Figure 3). Mean leaf inclination angle
was significantly greater for B. involucrata
compared to the other two species (ANOVA: F2,149
= 11.7, P < 0.0001; Tukey’s HSD: Piper sp.: P <
0.0001, R. spicata: P = 0.002).
The relationship between leaf-water shedding
efficiency and inclination angle was best
accounted for by fitting a quadratic regression for
Piper sp. and R. spicata (Piper sp: P = 0.02, r2 =
Table 1. Linear regression showed no correlation
between percent canopy cover and leaf
inclination angle for any species.
Statistical analyses
We used linear regressions to compare both
percent canopy cover and leaf height with leaf
inclination angle for each species. We conducted
an ANOVA with a Tukey’s HSD post-hoc
Species
Leaf inclination angle
A
slope
0.106 ± 18
0.82
0.001
R. spicata
-0.248 ± 19
0.57
0.007
B. involucrata
-0.166 ± 15
0.71
0.003
B
80
80
60
60
60
40
40
40
20
20
20
65
75
85
Percent canopy
cover
95
0
60
70
r2
Piper sp.
80
0
P
80
Percent canopy
cover
90
0
70
75
C
80
85
90
Percent canopy
cover
Figure 1: Leaf inclination angle was not correlated with percent canopy cover in any species (A:
Piper sp.; B: R. spicata; C: B. involucrata).
37
Leaf inclination angle
(°)
Palo Verde
A
80
B
80
60
60
60
40
40
40
20
20
20
0
40
110
0
180
Leaf height (cm)
60
C
80
110
160
Leaf height (cm)
0
20
70
120
Leaf height (cm)
Figure 2: Leaf inclination angle was not correlated with leaf height in any species (A: Piper sp.; B: R. spicata;
C: B. involucrata).
0.68; R. spicata: P = 0.01, r2 = 0.72) and a cubic
regression for B. involucrata (P=0.01, r2 = 0.86)
(Fig. 3).
DISCUSSION
Variability in light availability did not cause Piper
sp., R. spicata, or B. involucrata to modify leaf
angle to maximize exposure to solar radiation (Fig.
1). Similarly, there was no relationship between
leaf angle and leaf height in any species (Fig. 2),
suggesting that individual leaves within a plant do
not respond to variation in light availability.
The between-species difference in mean leaf
angle suggests that Piper sp., R. spicata, and B.
involucrata have leaf inclination angles suited to
the specific light and moisture conditions of their
preferred environments. As leaf inclination does
not vary in response to light for individual leaves
Table 2. Linear regression showed no
correlation between leaf height (cm) and leaf
inclination angle (°) for any species.
Species
Piper sp.
Slope
-0.040 ± 18
P
0.70
r2
0.003
R. spicata
-0.175 ± 19
0.23
0.03
B. involucrata
0.099 ± 15
0.42
0.01
100
Water shed (%)
95
Pip
er
sp.
Piper sp.
Poly.
(Piper)
R. spicata
Poly.
(Raz)
90
B. involucrata
Poly.
(Begonia)
85
80
-60
-40
-20
75
0
20
40
60
Leaf inclination angle (°)
Figure 3. Polynomial regressions of leaf water shedding efficiency by leaf inclination angle. Black shapes
are mean leaf inclination angle observed in the field for each species. Mean (±SE) leaf inclination angle
was -18.0° ± 3.9° for Piper sp., -23.6° ± 3.5 for R. spicata, and -39.6° ± 2.2 for B. involucrata.
38
Dartmouth Studies in Tropical Ecology 2014
or plants of any species, our results suggest that
the optimal inclination angle is a species-wide
adaptation allowing for specialization in a specific
microenvironment.
We found that B. involucrata, an early
colonizer species specialized to high-light
environments (McMillan et al. 2006), displays an
average leaf inclination angle near the
experimentally determined optimal range of water
shedding (Fig. 3). Piper sp. and R. spicata are not
primarily located in light-abundant disturbed areas
(Haber 2000) and therefore display morphologies
more suited to maximizing photosynthesis than
shedding water. Both species displayed mean leaf
angles far closer to horizontal than the nearvertical angles that are optimal for water shedding
(Fig. 3). However, it is possible that factors
besides leaf inclination angle, such as surface area
and drip tip length, also affect these species’
ability to tolerate or shed water (Meng et al. 2014).
Limited light availability and high moisture
levels in the tropical cloud forest strongly shape
the morphology and life history of understory
plant life (Lightbody 1985). Nutrient, light, and
moisture conditions vary spatially and temporally
due to disturbance, topography, aspect, and soil
substrate (Bazzaz and Pickett 1980). Tropical
understory plants have evolved diverse strategies
suited to highly specific environmental conditions
instead of adapting to tolerate and respond to a
broader range of conditions. The physiological
mechanisms through which these plants maximize
fitness in a range of microhabitats sheds light on
the underlying drivers of plant diversity in tropical
cloud forests.
LITERATURE CITED
Ackerly, D.D. and E.A. Bazzaz. 1995. Leaf
dynamics, self-shading and carbon gain in
seedlings of a tropical pioneer tree. Oecologia
101: 289-298.
Bazzaz, F. and S. Pickett. 1980. Physiological
ecology of tropical succession: a
comparative review. Annual Review of
Ecology and Systematics 11: 287-310.
Dean, J.M. and A.P. Smith. 1978. Behaviour and
morphological adaptations of a tropical
plant to high rainfall. Biotropica 10: 3441-47.
Falster, D.S. and M. Westoby. 2003. Leaf size and
angle vary widely across species: what
consequences for light interception? New
Phytologist 158: 509–525.
Haber, W.A. 2000. Vascular plants of
Monteverde. Pages 457-518 in N.M. Nadkarni
and N.T. Wheelwright, editors. Monteverde:
ecology and conservation of a tropical cloud
forest. Oxford University Press, New York.
Lightbody, J.P. 1985. Distribution of leaf shapes
of Piper sp. in tropical cloud forest:
evidence for the role of drip-tips. Biotropica
17: 339-342.
Lücking, R. and A. Bernecker-Lücking. 2005.
Drip-tips do not impair the development of
epiphyllous rain-forest lichen communities.
Journal of Tropical Ecology 21: 171-177.
McMillan, P.D., G. Wyatt, and R. Morris. 2006.
The Begonia of Veracruz: additions and
revisions. Acta Botánica Mexicana 75: 77-99.
Meng, F., R. Cao, D. Yang, K.J. Niklas, and S.
Sun. 2014. Trade‑ offs between light
interception and leaf water shedding: a
comparison of shade‑ and sun‑ adapted
species in a subtropical rainforest. Oecologia
174: 13–22.
Poorter, L. and D.M.A. Rozendaal. 2008. Leaf size
and leaf display of thirty-eight tropical tree
species. Oecologia 158: 35–46.
Rozendaal, D.M.A., V.H. Hurtado, and L. Poorter.
2006. Plasticity in leaf traits of 38 tropical tree
species in response to light; relationships with
light demand and adult stature. Ecology 20:
207-216.
ACKNOWLEDGEMENTS
We would like to thank J. Trout-Haney for her
guidance in developing methods, as well as the
Dartmouth Biology FSP 2014 for their assistance
in the peer review process.
AUTHOR CONTRIBUTIONS
C. Pendergrast and E. Plane collected field data
and M. DeGroote conducted all image analysis.
All authors contributed equally.
39
Palo Verde
FOREST STRUCTURE AND EPIPHYTE RICHNESS
NATHANAEL J. FRIDAY, LARS-OLAF HÖGER, ZACHARY T. WOOD
Faculty Editor: Ryan Calsbeek
Abstract: Trees play an important role in structuring forest habitats and climate. In cloud forests, many species of
epiphyte colonize trees. These epiphytes alter water and nutrient cycling. We examined the effects of forest stand
age on trunk epiphyte richness in the cloud forest of Monteverde, Costa Rica. We found that epiphyte richness per
trunk was positively influenced by tree diameter, and negatively influenced by stand age. Younger stands have
higher epiphyte richness due to higher densities of epiphytes per square meter of bark and greater overall stem
surface area. We demonstrate that as stands age, they may limit the richness of their co-occurring communities,
thereby decreasing the ecosystem processes these communities provide.
Key Words: Tropical forests, epiphytes, species richness, ecosystem functions
INTRODUCTION
The physical structures of trees can alter the
abiotic environment of a forest. Trees alter
available light, moisture, wind, and nutrient
regimes. Because forests undergo large changes
over the course of their succession, the way these
factors are influenced may change over time.
In cloud forests, trees also provide substrates
for epiphytic communities, which often occupy a
large portion of tree surfaces and facilitate a
variety of ecosystem processes. Because epiphytes
rely on trees as a substrate, the structural
characteristics of a tree stand can play a major role
in structuring epiphyte communities. Factors such
as elevation, light availability, and humidity have
all been shown to affect epiphyte richness (Holz et
al., 2002). Therefore succession can alter epiphyte
richness through tree morphology and microclimactic change. (Flores-Placios and GarciaFranco, 2006).
Because epiphytes are sensitive to the size,
shape, and local assemblage of forest trees,
epiphyte communities should change as a forest
develops. We therefore hypothesized that epiphyte
richness would be affected by successional age
and tree diameter.
In this paper, we will use the terms “young”
and “old” to refer to forests that are at an earlier or
later successional state. We do not intend to make
judgments on individual tree age.
METHODS
We haphazardly established 14 plots, each around
a focal tree of random size in the cloud forest of
Monteverde, Costa Rica. Within an eight-meter
40
radius of the focal tree, we measured diameter at
breast height (DBH, H = 1.37m) for all trees with
a DBH  5cm. For all trees with a DBH  5cm
within a five-meter radius of the focal tree, we
tallied the number of trunk epiphyte species within
a one meter vertical range centered at breast
height. We counted all woody stems with a DBH <
5cm and height  1.37 meters within the eightmeter radius.
We predicted epiphyte richness using a
regression including DBH, stem ratio (SR), and an
interaction between DBH and SR. SR is defined as
the ratio of woody plants with a DBH  5cm to
woody plants with a DBH < 5cm and H > 1.37m.
This ratio of large to small stems allows the
approximation of age structure in a forest stand,
with the ratio increasing as succession progresses
(Harper 601-602). We also fit a log-log regression
of epiphyte species density per square meter of
bark to DBH. Finally, we used a linear regression
to compare stand basal area, calculated as the sum
of the cross-sectional areas of all trees in a plot, to
SR.
RESULTS
The model we tested, in which all terms were
significant, included DBH, SR, and the interaction
between SR and DBH as the predictors of trunk
epiphyte richness. Trunk epiphyte richness
increased with DBH, but decreased with
increasing SR (Figure 1, Table 1). Greater SR also
decreased the slope of the richness ~ DBH line.
We found a negative linear relationship between
log-transformed epiphyte richness per square
Dartmouth Studies in Tropical Ecology 2014
meter of bark and log-transformed DBH (Figure 2;
F3,170 = 68.54, P < 0.0001). Basal area did not
change with SR (F1,12 = 0.0485, P = 0.83).
Figure 1. Epiphyte richness per trunk increases with
diameter at breast height (DBH). The lower dashed line
is the richness~DBH relationship when the stem ratio
(SR; large/small) is at mean + one standard deviation,
and the upper is when the SR is at mean – one
standard deviation (R2 = 0.54).
Figure 2. Species density per square meter bark
decreases with tree diameter (R2 = 0.69.). Both axes are
on a log-scale.
Table 1. Regression results for epiphyte richness.
Effect
Est.
Std. Err.
t
p
Intercept
DBH
SR
DBH*SR
0.289
0.006
0.555
0.031
16.7
12.1
-4.19
-3.44
<0.00001
<0.00001
<0.00001
0.00007
4.83
0.067
-2.32
-0.107
DISCUSSION
Our results supported the hypothesis that stand
age, as well as individual stem diameter, affects
epiphyte richness. Although the number of
epiphyte species on a tree increased with the
diameter of a tree (Figure 1), epiphyte richness per
square meter declined with increasing diameter
and total surface area (Figure 2).
Our results show that younger stands have a
greater richness of trunk epiphyte species. Stand
basal area in our study forest remains relatively
constant as stands age and average tree size
increases. Younger stands therefore have a low SR
and high overall stem surface area. The
combination of greater species richness per square
meter (Figure 2) and greater total surface area
cause younger stands to have a greater epiphyte
species richness overall.
We conclude that as stands age, epiphyte
species richness will decrease. Because a high
richness of plants is necessary to maintain
ecosystem processes (Isbell et al., 2011),
decreasing epiphyte richness in older forests may
thereby reduce the ecosystem processes
performed. These processes include atmospheric
nitrogen retention, arboreal sedimentation (Clark
et al., 2005), rainfall interception (Hölscher et al.,
2004), and habitat provision for inquiline
communities (Richardson et al., 2000, Armbruster
et al., 2002).
41
Palo Verde
Our results indicate that trees play an
important role in shaping epiphytic communities
and altering the ecosystem processes made
available by such communities. We demonstrate
that as forests age, they may limit the richness of
their surrounding communities and thereby the
quality of the ecosystem processes these
communities provide.
ACKNOWLEDGEMENTS
We would like to thank the Dartmouth 2014 FSP
group for their feedback and reviewer comments.
We would also like to thank the FSP faculty for
their assistance in editing this manuscript.
LITERATURE CITED
Armbruster, P., Hutchinson, R. A. & Cotgreave, P.
2002. Factors influencing community structure
in a South American tank bromeliad fauna.
Oikos, 96: 225–234.
Cardelus, C. L., Colwell, R. K. and Watkins, J. E.,
2006. Vascular epiphyte distribution patterns:
explaining the mid-elevation richness peak.
Journal of Ecology, 94: 144–156.
Clark, K.L., Nadkarni, N.M., Gholz H.L., 2005.
Retention of inorganic nitrogen by epiphytic
bryophytes in a tropical montane forest.
Biotropica. 37: 328-336.
Flores-Palacios, A. and García-Franco, J. G.,
2006. The relationship between tree size and
epiphyte species richness: testing four
different hypotheses. Journal of
Biogeography, 33: 323–330.
Harper, J.L., 1977. Population Biology of Plants.
Academic Press.
Holz, I., Robbert-Gradstein, S., Heinrichs, J.,
Kappelle, M., 2002. Bryophyte diversity,
microhabitat differentiation, and distribution
of life forms in Costa Rican upper montane
quercus forest. The Bryologist. 105: 334-348.
Hölscher, D., Köhler, L., Van Dijk, A.I.J.M.,
Bruijnzeel, L.A., 2004. The importance of
epiphytes to total rainfall interception by a
tropical montane rain forest in Costa Rica.
Journal of Hydrology. 292: 308-322.
Isbell, F., Calcagno, V., Hector, A., Connolly, J.,
Sanley-Harpole, W., Reich, P.B., SchererLorenzen, M., Schmid, B., Tilman, D., Van
Ruijven, J., Weigelt, A., Wilsey, B.J.,
Zavaleta, E.S., Loreau, M., 2011. High plant
diversity is needed to maintain ecosystem
services. Nature.477: 199-202.
Richardson, B.A., Richardson, M.J., Scatena, F.N.,
McDowell, W.H., 2000. Effects of nutrient
availability and other elevational changes on
bromeliad populations and their invertebrate
communities in a humid tropical forest in
Puerto Rico. Journal of Tropical Ecology.
16:167-188.
42
AUTHOR CONTRIBUTIONS
All authors contributed equally.
Dartmouth Studies in Tropical Ecology 2014
DIFFERENTIAL MOBBING IN TROPICAL HIGHLAND BIRDS
ANN M. FAGAN, EMILY R. GOODWIN, AND AARON J. MONDSHINE
Faculty Editor: Matthew P. Ayres
Abstract: Some animal species display context-specific defensive behaviors, which presumably provide potential
benefits while minimizing energy expenditure. Predator mobbing is an action in which groups of animals such as
birds incur high energetic costs and assume predation risk by flying directly at a predator to deter it. Therefore, birds
should selectively mob predators that pose a large risk. We tested the hypotheses that bird response level would vary
with both predator type and size class of bird. In the tropical highlands at Cuerici Biological Reserve in Costa Rica,
we played the calls of three local owls - the spectacled owl (Pulsatrix perspicillata), the mottled owl (Strix virgata),
and the Andean pygmy owl (Glaucidum jardinii) - and counted the number of extra-small, small, and medium sized
birds that responded along our transect. The Andean pygmy owl call induced the highest response level by birds,
and smaller birds displayed stronger responses than larger birds. This variation in response level suggests that the
energetic costs of mobbing are not equal for birds of different sizes. Our results indicate that mobbing response is
influenced by bird size and predator type. The behavior could involve a combination of innate predator recognition
and learning.
Keywords: avian behavior, cooperative defense, mobbing
INTRODUCTION
Defense against predators is a necessary energetic
cost for many species (Lima and Dill 1990).
Whether organisms use crypsis, toxicity, mimicry,
or other defensive strategies, energy spent on
defense cannot be used for other vital functions
such as reproduction or foraging. Selection
pressure to complete these other tasks drives
organisms to reduce energy spent on defense while
minimizing fitness costs. This pressure could drive
the evolution of context-specific defensive
behaviors.
Mobbing is a defensive behavior exhibited by
many bird species. Detection of a nearby predator
sometimes prompts groups of birds to harass the
predator with vocalizations and physical attacks.
This behavior benefits all members of the group,
even those not who are not being directly attacked,
by driving the predator away from the area and
possibly reducing the probability that the predator
will return. However, mobbing requires costly,
rapid flight as well as increased predation risk in
flying directly at the predator. Because of the high
cost and risk involved, birds are expected to only
exhibit this behavior in response to a perceived
threat.
In addition, different bird species have
evolved to recognize and respond to unique sets of
threatening stimuli. In some bird species, juvenile
birds learn which predators to mob by observing
adult birds responding to predator threat. Thus, an
adult bird’s failure to respond to a predator call
may be due to one of two things: first, that the bird
has never heard the call before and does not
associate it with predator threat, or second, that the
bird has heard the call, but that the predator is not
sufficiently threatening to prompt a mobbing
response. Because birds only mob certain
predators, we expected that birds of Cuerici
Biological Reserve would vary in the degree of
mobbing depending on the species of predator.
Bird species within the same range may also
respond differently to the same predator call. This
disparity may be due to the fact that the costs and
benefits of mobbing are not shared equally by
birds across different size classes. A small bird
may be more easily depredated by an aerial
predator and thus have more to gain from
engaging in cooperative defense. In this case,
smaller birds would respond more strongly to a
predator call. Alternatively, larger birds may be
able to more easily assume the high energetic cost
required to sustain the costly flight patterns of
mobbing. If this is true, larger birds should
respond most strongly to hearing a predator call.
METHODS
Study System
Our study focused on the mobbing behavior of
tropical highland birds at Cuerici Biological
43
Cuericí
Data Collection
We conducted 19 trials between the hours of
05:30 and 18:00 on 29-31 January 2014 in Cuerici
Biological Reserve. Our two study sites were a
partially covered brush field and a small stretch of
secondary forest trail. We began each trial with a
control period of three minutes, during which we
recorded number and size class of all birds seen
(stationary or moving) within a 3-meter radius of
our transect area. We classified birds with a body
length of <10 cm as extra-small, 10-15 cm as
small, and >15 cm as medium. After completing
the control period and waiting one minute, we
began the three-minute treatment period by
playing a randomly selected 90-second recording
of either a Spectacled owl, Mottled owl, or an
Andean pygmy owl followed by 90 seconds of
silence. During the treatment period, we again
recorded all birds seen within the specified radius
of the trial area.
Statistical Analyses
We evaluated a generalized linear model with
predator type and bird size as independent
variables and overall bird response as the response
variable. Bird response was defined as the
difference in number and size of birds present
44
during treatment and control trials. We found no
effect of time of day (morning vs. afternoon) (t =
0.30, P = 0.76, df = 38) or location (secondary
forest vs. brush field) (t = 0.72, P = 0.48, df = 38)
on bird response. All analyses were performed
using JMP® 11.0 statistical software (SAS
Institute, Cary, NC).
RESULTS
We observed 253 birds of at least eight different
species during our study. The extra-small species
included volcano hummingbirds (Selasphorus
flammula) and green violet-ear hummingbirds
(Colibri thalassinus). The small species included
black-and-white becards (Pachyramphus
albogriseus), black-cheeked warblers
(Basileuterus melanogenys), collared redstarts
(Myioborus torquatus), rufous-naped wrens
(Campylorhynchus rufinucha), Wilson’s warblers
(Cardellina pusilla), and some unidentified
species. The medium species were large-footed
finches (Pezopetes capitalis) and various
unidentified flycatcher species (Ptiliogonatidae
and Tyrannidae).
The Andean pygmy call induced significantly
higher total bird response than the mottled or
spectacled owl calls (F2,111 = 8.37, P = 0.0004; Fig
1., Table 1) There was a significant effect of the
interaction between predator and bird size (F4,111 =
2.56, P = 0.043) with smaller birds exhibiting
higher levels of response than larger birds.
Mean Response
Reserve, Costa Rica (elevation 2500 meters) in
response to three avian predators found within
Costa Rica: the spectacled owl Pulsatrix
perspicillata, the mottled owl Strix virgata, and
the Costa Rican pygmy owl Glaucidium
costaricanum. Although the Costa Rican pygmy
owl was taxonomically separated from the Andean
pygmy owl (G. jardinii) in 2000, both owls have
the same vocal repertoire and therefore we
substituted the call of the Andean pygmy owl for
the call of the Costa Rican pygmy owl. Of the
three owls used, P. perspicillata and S. virgata are
not likely to be found at this high elevation. All
three owls occasionally feed on small birds,
although P. perspicillata and especially S. virgata
depend more on small mammals, insects, lizards,
and frogs for food. G. jardinii is the smallest of the
three owls (average body length 15 cm), followed
by the mottled (36 cm) and spectacled (48 cm)
owls (Stills & Skutch 1989). All three species
have been reported to be mobbed in the daytime
by small birds (Stills & Skutch 1989, Cornell
Neotropical Birds database).
3
2.5
2
1.5
1
0.5
0
-0.5
-1
Andean
Mottled
Spectacled
XS
S
Bird Size
M
Figure 1. Total bird response to the Andean pygmy
owl call was higher than to the other two owl calls.
Within the Andean pygmy owl trials, medium birds
were less responsive than smaller birds. Mean
response was defined as the difference in mean
number of birds of each size class between treatment
and controls.
Dartmouth Studies in Tropical Ecology 2014
Table 1. ANOVA of the effects of predator and bird size on level of bird response.
DF
Mean Squares
F
P
Predator
2
26.43
8.37
0.0004
Bird Size
2
2.785
0.88
0.42
Predator*Bird Size
4
8.065
2.56
0.043
111
3.157
Error
DISCUSSION
Cooperative defense can be costly and dangerous.
Thus, prudent birds are expected to mob only in
situations where the predator represents a potential
threat. In our study, birds responded only to the
Andean pygmy owl call, indicating bird specificity
of the mobbing response. Of the three predators,
only the Andean pygmy owl occupies the
geographic and elevational range in which we
performed our experiment. Since the mottled and
spectacled owls are not found at the high elevation
of our study area, the birds may not have
recognized the calls of P. perspicillata or S.
virgata. Alternatively, the increased individual
risk associated with mobbing the two larger owl
species could outweigh the potential benefits of
group defense.
Birds of different size classes also responded
differentially to the Andean pygmy owl call.
Smaller birds were more responsive than larger
birds, which was consistent with the hypothesis
that small birds have more to gain from mobbing
behaviors. It seems likely that larger birds are less
likely to be depredated by a small owl like the
Andean pygmy owl, in which case the small
predation risk would not justify the energetic costs
of mobbing.
The results of our study suggest that avian
defensive behaviors in the tropics are contextspecific. Not only do birds respond to certain
threats and not others, but their level of response
may vary by size.
We were unable to resolve the extent to which
responses were learned versus innate. Indeed,
innate recognition of predators and learned antipredatory behaviors are not mutually exclusive.
Additional studies would be required to
distinguish among the possibilities. The outcome
would have particular relevance to newlyintroduced or invasive avian predators and their
effects on local bird communities. For example,
birds that only have innate predator recognition
may be at a disadvantage compared to those that
learn from a wide variety of stimuli, as the former
would not be able to recognize the call of a newlyintroduced avian predator. Additionally,
introduced predators may have significant effects
on community structure, as juvenile birds may
have an advantage over adult birds in learning to
recognize new predator calls. Beyond differences
in size class, relative dependence on innate
predator recognition versus learned anti-predator
defenses may determine the outcome of birdpredator interactions. Regardless of the
mechanism by which birds recognize a predator
call, the benefits of group defense presumably
outweigh the costs of mobbing or the behavior
would not be so well-developed.
LITERATURE CITED
Caro, T. M. Antipredator Defenses in Birds and
Mammals. Chicago: University of Chicago,
2005.
Sandoval, L., and D.R. Wilson. 2012. Local
predation pressure predicts the strength of
mobbing responses in tropical birds. Current
Zoology 58: 781–790.
Stiles, F.G., and A.F. Skutch. A Guide to the Birds
of Costa Rica. Ithaca, NY: Comstock, 1989.
Lima, S.L. and L.M. Dill. 1990. Behavioral
ACKNOWLEDGEMENTS
We thank Carlos Solano for the use of the grounds
at Cuerici Biological Reserve. We are also
indebted to Jessica Trout-Haney for her assistance
in conducting trials and identifying birds.
AUTHOR CONTRIBUTIONS
All authors contributed equally.
45
Cuericí
decisions made under the risk of predation: a
review and prospectus. Canadian Journal of
Zoology 68: 619-640.
46
The Cornell Lab of Ornithology: Neotropical
Birds. Cornell Lab of Ornithology. Web.
Accessed 09 Feb. 2014. http://neotropical
.birds.cornell.edu.
Dartmouth Studies in Tropical Ecology 2014
DIVERSITY OR DOMINANCE: SYNCHRONOUS MASTING AS A REPRODUCTIVE STRATEGY
IN QUERCUS COSTARICENSIS MONTANE FORESTS
MAYA H. DEGROOTE AND CLAIRE B. PENDERGRAST
Faculty Editor: Matthew P. Ayres
Abstract: Montane oak forests represent a unique, biologically rich ecotype in the American Tropics. These forests
support distinctive plant and animal communities suited to the biotic and abiotic conditions of an oak-dominated
forest, a rare occurrence in a region otherwise characterized by exceptionally high tree species richness. Several
theories offer possible explanations for the high diversity and heterogeneity of tropical forests. The Janzen-Connell
model identifies specialist predation near parent trees as the primary driver of seedling distribution in tropical
forests, while the predator-swamping model asserts that generalist predators are primarily responsible for controlling
seedling establishment. We assessed the relative capacity of these two forces to explain patterns of seed distribution
and viability in the montane oak forest of Cuerici, Costa Rica. We surveyed acorn abundance, predation type, and
viability at high- and low-density plots in the interior and at the edge of a masting Quercus costaricensis stand. We
found that a higher percentage of acorns were viable at low-density plots, with greater specialist insect predation in
interior areas and greater generalist vertebrate predation at the edge of the masting stand. Acorn viability per plot
was consistent across locations and densities, suggesting that the combined impact of specialist and generalist
predation across a masting area allows for evenly distributed establishment of Q. costaricensis seedlings and a
continuation of oak-dominated canopy cover. These findings are relevant to understanding how disturbance affects
seedling establishment in tropical oak forests and how deforestation in these unique environments alters ecosystem
structure and species diversity.
Key words: acorns, Janzen-Connell hypothesis, predator-swamping, Quercus costaricensis, synchronous masting
INTRODUCTION
Tropical forests harbor hyper-diverse plant
communities, far exceeding diversity levels found
outside the tropics (Wright 2002). The JanzenConnell hypothesis attributes high species richness
of tropical trees to the combination of specialist
seed predators and seed dispersal limitations.
Under this model, specialist predators limit seed
success in high seed density areas close to the
parent tree, resulting in greater likelihood of
seedling establishment in areas farther from the
parent tree at the outer edges of the dispersal range
where specialist predation is less prevalent (Janzen
1970). Thus, conspecific neighboring trees are
rare, resulting in high species richness.
The Janzen-Connell model suggests that
seedling survival would be chronically low in low
diversity forests due to high specialist seed
predator abundance (Janzen 1970). However,
some trees employ synchronous masting, a
reproductive strategy in which neighboring
conspecific trees simultaneously produce a high
quantity of seeds, while in other years few or no
seeds are produced. Tree species that mast
synchronously may be able to establish seedlings
by overwhelming the capacities of predators to eat
all the seeds produced, a phenomenon known as
predator-swamping (Ims 1990).
While the Janzen-Connell hypothesis predicts
lower seed survivorship at higher seed density due
to the pressure of specialist predators, predatorswamping theory predicts higher survivorship
when an area is inundated with a high density of
seeds. Near the Cuerici Biological Reserve in
Costa Rica, a single tree species, Quercus
costaricensis, dominates the high altitude
mountain tops. Q. costaricensis mast
synchronously every 3-4 years and are able to
sustain an apparently stable forest type in which
nearly 100% of the canopy trees are a single
species (Solano pers comm, Guariguata and Sáenz
2002). Q. costaricensis timber is valuable and
many of Costa Rica’s oak-dominated forests have
experienced increased pressure as a result of
expanded logging activity. The resilience of these
forests depends largely on Q. Costaricensis’
ability to successfully establish seedlings and
maintain dominance. Loss of oak forest would
affect ecosystem structure in Cuerici, with broad
consequences for the many organisms dependent
on oaks for food and habitat.
47
Cuericí
In this oak-dominated forest, insects are hostspecific predators while vertebrates are typically
generalist seed predators (Kellner et al. 2009). We
tested whether the Janzen-Connell model or
predator-swamping more strongly determined Q.
costaricensis seed survival and subsequent
establishment. We surveyed low- and high-density
acorn plots within and on the edge of the masting
area for Q. costaricensis in 2013-2014. The
Janzen-Connell model predicted highest seed
mortality in high-density interior plots, where
specialist predators would be most abundant,
while the predator-swamping model predicted
higher seed mortality in low-density edge areas
where predators are less likely to be satiated.
METHODS
On 31 January, 2014, we surveyed one square
meter plots along a cleared 0.5 km trail through a
masting stand of Q. costaricensis centered at N
9.56443° and W 83.66628° at 2750 meters
elevation (Fig. 1). We identified the edge of the
masting area and sampled twelve plots at this
edge, in the seed shadow (Janzen 1970) of the last
masting tree. Six of these plots were deliberately
located where there was a high density of acorns
and six where there was a low density of acorns in
the plot. We similarly sampled 14 plots within the
interior of the masting area (approximately 100
meters from the edge plots). Seven of these
interior plots were deliberately located where
acorn density would be high (beneath the crowns
of three masting trees) and seven were located
where density would be low (below inter-tree gaps
in the canopy). All acorns within a plot were
collected and categorized as either viable or killed,
with the causes of mortality being noted as insect,
vertebrate, or other (typically fungi).
RESULTS
Acorn density varied dramatically between plots,
ranging from 5 to 114 acorns per square meter.
Dominant predation type differed significantly
between interior and edge plots. Percent mortality
due to vertebrate predation was higher in edge
plots (Fig. 2A). High-density edge plots displayed
the greatest percent mortality from vertebrate
predation (72.3 ± 3.8%). Interior plots suffered
significantly higher percent mortality from insect
predation than edge plots (Fig. 2B). Insect
predation was greatest in interior high-density
plots (46.3 ± 8.3%), while edge low-density plots
experienced the lowest mortality from insect
predation (10.7 ± 6.1%).
Interior
Edge
≈ 60m
≈ 2.5m
≈ 10m
Low-density plot
High-density plot
≈ 50m
Q. costaricensis
Figure 1. Location of high-density and low-density plots along a cleared trail in the interior and edge of the
masting stand. Dotted lines indicate tree canopy.
48
Dartmouth Studies in Tropical Ecology 2014
Mortality from
vertebrate predation
(%)
A
80
low density
70
high density
60
50
40
30
20
10
0
B
50
40
(%)
Mortality from
insect predation
60
30
20
10
0
Edge
Interior
Location
Figure 2. (A) Percent mortality from vertebrate
predation was significantly higher in edge plots
than in interior plots (t = 3.23, P = 0.004, df =
19.3). There was no difference in percent mortality
from vertebrate predation between high- and lowdensity plots (t = 0.78, P = 0.45, df = 16.1). (B)
Percent mortality from insect predation was
significantly higher in interior plots than in edge
plots (t = 2.67, P = 0.01, df = 20.5). Percent
mortality from insect predation was significantly
higher in high-density than low-density plots (t =
2.5, P = 0.02, df = 17)
Insect-depredated acorns all featured a round hole
approximately one millimeter in diameter. When
cut open, these seeds displayed darker regions
extending towards the center of the seed. We
assumed these seeds to be non-viable due to
predation by specialist acorn-feeding insects,
perhaps weevils (Kellner et al. 2009). Vertebratedepredated acorns displayed ragged holes roughly
two to six millimeters wide and endosperm was
often entirely absent. During our sampling period
we observed acorn woodpeckers and variegated
squirrels in the canopy of masting trees. These
same species were observed in non-masting stands
as well, though they appeared to be most abundant
in the masting area.
We observed consistently higher acorn
densities in the interior than in the edge for both
low-density and high-density plots (Fig. 3A).
Percentage of viable acorns was significantly
higher in low-density plots in both edge and
interior plot types (Fig. 3B). We found no
significant relationship between viable acorn
frequency and either plot density or plot location (t
= 1.82, P = 0.087, df = 17.0). However, highdensity interior plots contained more than twice as
many viable acorns (9.14 ± 2.9 acorns/m2) as any
other plot type (Fig. 3C).
DISCUSSION
Distribution and abundance of viable Q.
costaricensis acorns were determined by seed
dispersal from the parent tree as well as insect and
vertebrate seed predation. Patches of varying seed
densities existed within and at the edge of the
masting area due to the overlapping seed shadows
of neighboring masting trees (Fig. 2A). The
heterogeneous distribution of acorns generated an
array of microhabitats suited to different predation
strategies (Pérez-Ramos 2008). Vertebrate
predators with larger ranges may have a greater
impact on low-density areas on the edge of the
masting stand while less mobile insects dominate
high-density patches directly beneath masting
trees. Predators consumed the vast majority of
acorns produced by masting oaks. The
mechanisms allowing for continued oak
dominance in the Cuerici forest depend largely on
the success and distribution of different predator
types.
We found a significantly higher percentage of
viable acorns in low-density than high-density
plots (Fig. 3B). That is, seed predation impacted
high-density areas more strongly than low-density
areas, presumably because mobile predators tend
to forage preferentially in places with high prey
concentrations (Ims 1990). Our observations of
high predation in high-density areas due to the
abundance of seeds support both the JanzenConnell and the predator-swamping models of
seed predation. However, the two models assume
different predator groups as primarily responsible
for controlling seed viability and long-term
seedling establishment. The Janzen-Connell model
identifies specialist predators found beneath parent
trees as the primary threat to seed viability (Janzen
1970), while the predator-swamping model tends
49
Cuericí
to emphasize the importance of larger, more
generalized and more mobile seed predators (Ims
1990).
Vertebrate predation contributed most strongly
to acorn mortality in edge plots and was less
pronounced in interior plots (Fig. 2A). This is in
accordance with the predator-swamping model, as
generalist predators tend to consume prey
resources until they reach the satiation point
determined by their population size and prey
handling efficiency. Generalist predators can be
unable to consume all available seeds in high preydensity areas such as the interior of a synchronous
masting stand, resulting in an increased likelihood
of seedling establishment. Masting events are too
brief to allow vertebrate predator populations to
increase their satiety thresholds through
population growth or improved prey capture
strategies. Therefore, generalist predation favors
seedling establishment in interior areas where
seeds are most abundant.
Insect predators contributed more strongly to
acorn mortality in interior than edge plots (Fig.
2B). Such specialist predators display limited
dispersal capability and are therefore dominant in
high-density interior areas (Fig. 2B). Our findings
supported the prediction of the Janzen-Connell
model of high insect predation near the parent tree.
Percent mortality due to insects was nearly four
times greater in interior high-density plots than in
any other plot type (Fig. 2B).
How do the competing Janzen-Connell and
predator-swamping models of seed predation
explain the sustained dominance of Q.
costaricensis in the upper montane forests of
Costa Rica? Seedling establishment is dependent
on density of viable acorns, which we found did
not differ significantly with plot location or acorn
density (Fig. 3C). No single plot type offered
superior conditions for acorn survival due to the
complex impacts of both generalist and specialist
seed predators and dispersal limitations. While the
vast majority of acorns produced during a masting
event fail to become established (Fig. 3B), Q.
costaricensis’ synchronous masting strategy
evidently allows enough seedling establishment to
maintain oak canopy dominance. This is possible
because synchronous reproduction creates a range
of densities within the masting oak stand, allowing
acorns to evade specialist predation in highdensity areas. Synchronous masting also facilitates
generalist predator satiation by concentrating seed
production in time and space such that predator
population size is limited.
Once widely distributed throughout the
highland Neotropics, few intact blocks of montane
oak forest remain (Kappelle 2006). Patches of
primary oak forest as found in Cuerici, Costa Rica
face increasing fragmentation and disturbance due
to logging for timber and charcoal, as well as
clearcutting for agriculture or road construction.
The clearing of Neotropical montane forests has
led to land degradation and biodiversity depletion.
Interest in the protection and recovery of these
valuable forests has prompted many questions
regarding how forests should be managed and
monitored to reestablish stable, healthy forest
ecosystems (Guariguata and Sáenz 2002).
Recovery of tropical montane oak forests
following disturbance depends strongly on acorn
availability and activity of seed predators and
dispersers (Kappelle 2006). Our findings may
inform future monitoring and management efforts
to protect the biodiversity and ecosystem services
offered by tropical montane oak forests.
LITERATURE CITED
Guariguata, M. and G. Sáenz. 2002. Post-logging
acorn production and oak regeneration a
tropical montane forest, Costa Rica. Forest
Ecology and Management 167: 285-293.
Ims, R. 1990. Reproductive synchrony as a
predator-swamping strategy. The American
Naturalist 136: 485-498.
Janzen, D. 1970. Herbivores and the number of
tree species in tropical forests. The American
Naturalist 104: 501-528.
Kappelle, M. Ecolo gy and Conservation of
50
ACKNOWLEDGEMENTS
We would like to thank Vivek V. Venkataraman
and Amelia K. Ritger for their discovery of the
masting stand of Q. costaricensis and Matthew P.
Ayres for sparking our interest in the masting
system. Thank you also to Carlos Solano for his
hospitality and knowledge of natural history.
AUTHOR CONTRIBUTIONS
Both authors contributed equally
Dartmouth Studies in Tropical Ecology 2014
Neotropical Montane Oak Forests. Springer,
2006.
Kellner, K., J. Riegel, N. Lichti, and R. Swihart.
2009. Oak mast production and animal
impacts on acorn survival in the central
hardwoods. Gen. Tech Rep. 108: 176-187.
Pérez-Ramos, I. 2008. Factors affecting postdispersal seed predation in two coexisting oak
species: Microhabitat, burial and exclusion of
large herbivores. Forest Ecology and
Management 255: 3506-3516.
Solano, Carlos. Personal communication, January
2014.
Wright, J. 2002. Plant diversity in tropical forests:
a review of mechanisms of species
coexistence. Oecologia 130: 1-13.
51
Cuericí
INSECTIVOROUS BATS ALTER FORAGING DECISIONS BASED ON RISK OF VISUAL
PREDATION
ALEXA E. KRUPP, REBECCA C. NOVELLO, AND ELLEN R. PLANE
Faculty Editor: Matthew P. Ayres
Abstract: Prudent animals may adjust foraging behavior to limit predation threat. Nocturnal foragers are under
greater threat from visual predators when ambient light is relatively high. We tested whether insectivorous bats
would adjust their behavior in response to cues of visual predators in combination with the addition of light. We
recorded bat activity with an ultrasonic microphone under factorial combinations of light and simulated owl calls.
Bat passes were less frequent in trials with simulated predator calls, and with the experimental addition of light
without predator calls. These results indicate that an increase in risk of predation from auditory and abiotic cues,
including exposure to light, may induce predator avoidance behavior. Over time, selective pressures can reward
nuanced predator detection and avoidance behaviors, even at the cost of decreased energy intake from foraging.
Keywords: bats, optimal foraging theory, visual predators
INTRODUCTION
Optimal foraging theory predicts that animals
should adopt foraging strategies that maximize
their fitness via high resource capture (Pyke et al.
1977). However, risk of predation forces prey
animals to balance foraging behaviors with
predator detection and avoidance (Kavaliers and
Choleris 2001). Bats, the only true mammalian
fliers, have extremely high metabolic rates during
active flight, 3-5 times higher than the maximum
of equivalent-sized terrestrial mammals (Shen et al
2010). When foraging at night, insectivorous bats
can fly continuously for up to seven hours,
locating and capturing prey using high-frequency
echolocation. This high energetic expenditure is
fueled through rapid metabolism of ingested
nutrients (Voigt et al. 2010).
As nocturnal foragers with visual predators,
bats are more vulnerable to predation in higher
ambient light (Lima and Dill 1990). Many
insectivorous bats remain in their roosts during the
high insect activity at dusk, presumably to avoid
higher predation risk before full darkness (Rydell
1996). Similarly, some frugivorous bats avoid
foraging during the full moon (Clarke 1983), and
nocturnal rodents reduce foraging under
experimentally manipulated high light conditions
(Kotler 1984). On the other hand, light-induced
reductions in foraging would needlessly limit
resource capture if predators are absent.
We tested whether insectivorous bats in the
montane forests of Costa Rica adjusted their
foraging behavior in response to different
52
combinations of light and simulated predator risk.
Specifically, we evaluated effects of predator calls
and manipulation of light on the frequency and
duration of bat passes, which are both proxies for
foraging activity. If bats are responsive to variable
predation risks, they should reduce foraging in
response to signals of predators, especially with
increased light levels. Alternatively, if predation
has been only a minor factor in bat fitness, or at
least less important than meeting metabolic needs,
bats may be nonresponsive to light or cues of
predator presence.
METHODS
We recorded bat activity at Cuericí Biological
Station on the nights of 29-30 January, 2014. We
used Avisoft Bioacoustics hardware and software
(Avisoft Bioacoustics, Berlin, Germany) and an
ultrasonic microphone to record bat calls between
the hours of 18:30 and 21:00. In 40-minute
periods, we conducted four ten-minute trials,
randomly assigning treatments to time slots within
this period. We manipulated the presence and
absence of light and the presence and absence of
predator calls to create four treatments. We turned
on two lights over the lower Cuericí trout pond for
light trials. To simulate predator calls, we played
recordings of an Andean Pygmy Owl and a Striped
Owl for 30 seconds every 90 seconds during the
trial. Both the Andean Pygmy Owl and Striped
Owl are found at high elevation sites in and
around Cuericí and are known to prey on bats and
other small mammals (Stiles and Skutch 1989).
Dartmouth Studies in Tropical Ecology 2014
We used SASLab Pro (Avisoft Bioacoustics,
Berlin, Germany) to analyze recordings for bat
passes. We recorded passes as independent if there
was at least one second of silence between calls.
We also randomly sampled ten bat passes from
each trial to assess bat pass duration. We used bat
passes and pass durations to capture changes in
two potentially important measures of foraging
behavior: the frequency with which a bat visits a
foraging area (pass) and the time it spends in the
area on each visit (pass duration).
Table 1. A log-linear model showed significant
effects of time of night (period) and predation on
number of bat calls, as well as an interaction
between predation and light.
Statistical Analysis
We performed a chi-square analysis to test for
effects of light and predator calls on bat passes for
each night. We also tested for effects of light,
predator calls, period, and interactions of the three
factors on bat passes for each night using a three
way log-linear model (Quinn and Keough 2002).
We performed an ANOVA to compare bat pass
durations across treatments. We conducted all
analyses in JMP 10 (SAS Institute, Cary, NC).
1), including a Myotis-type call (Fig. 1a) and a
Phyllostomidae-type call (Fig. 1c). We saw
evidence of search calls, approach calls, and
feeding buzzes (Fig. 2).
Frequency of bat passes was greater on the
first night (N=178) than the second night (N = 44).
On the first night, we found a significant
difference in frequency of bat passes across
treatment types (chi-square = 57.96, P < 0.0001, df
= 1; Fig. 3). We observed more bat passes in the
absence of predator calls, and number of bat
passes varied with both period and the interaction
between light and predator calls (Table 1). The
most passes were observed during dark treatments
without predator calls (Fig. 3).
RESULTS
We recorded 222 bat passes in 73 sound files
(Electronic supplementary materials). We
observed three different bat call sonotypes (Fig.
P
19.75
<0.0001*
Predation
12.45
0.0004*
Light
0.03
0.87
Predation*Light
35.63
<0.0001*
Period*Light
0.61
0.43
Period*Predation
0.20
0.65
B
C
Frequency (kHz)
A
Chi-square
Period
Frequency (kHz)
Figure 1. Sonograms of three bat searching calls recorded over 20 milliseconds. A) Steep frequencymodulated sweep with drop off in frequency at tail end of call (Myotis-type) B) Relatively narrow bandwidth
at end of call C) Relatively short call with three harmonic frequencies (Phyllostomidae-type).
Figure 2. Bat feeding buzz with steep frequency-modulated sweep, recorded over 260 milliseconds.
53
Cuericí
No factor had a significant effect on bat passes on
the second night, but when we combined data
from both nights the effects of period (chi-square
= 15.59, P < 0.0001 df = 1), predation (chi-square
= 8.52, P = 0.004, df = 1), and interaction between
light and predator calls (chi-square =19.54, P <
0.001, df = 1) remained significant.
Duration of calls ranged from 0.5 to 16.7
seconds (mean ± SE = 5.25 ± 3.6 sec), and neither
predator calls, light, nor the interaction between
light and predator calls had a significant effect on
call duration (F1,36 < 1, P > 0.65 for all factors).
Figure 3. Bat passes on the first night were most
frequent in the dark without predator calls. See
Table 1 for corresponding analyses.
DISCUSSION
We found that bats adjusted their foraging
behavior in response to increased risk from
visual predators. Bat passes were markedly less
frequent in the presence of predator calls on the
first night of our study (Table 1, Fig. 3),
indicating that bats reduce foraging efforts when
risk of predation is high. Light alone did not
have a significant effect on bat passes on the
first night, but the effect from interaction of
light and predator calls was significant (Table
LITERATURE CITED
Clarke, J.A. 1983. Moonlight’s influence on
predator/prey interactions between short-eared
owls (Asio flammeus) and deermice
(Peromyscus maniculatus). Behavioral
Ecology and Sociobiology 13: 205-209.
Kavaliers, M., and E. Choleris. 2001. Antipredator
responses and defensive behavior: ecological
and ethological approaches for the neurosci-
54
1). In the absence of predator calls, bat passes
were less frequent in the light, where risk of
visual predation is greater (Fig.3). Thus, bats
seemed to exhibit predator avoidance in
response to both light levels and auditory cues.
This supports the hypothesis that bats can adjust
their behavior in nuanced ways according to
details of predation threat (e.g., the combination
of light with the call of a predator that hunts by
sight).
Neither light nor predator calls had an effect
on bat passes on the second night, which could
indicate bat habituation to our experimental
stimuli. Further studies could investigate
whether bats respond differentially to predation
threat based on the relative abundance of
predators in their environment, which would
suggest the extent to which predator avoidance
behavior in bats is innate versus learned.
Our findings provide evidence for a tradeoff
between foraging and predator avoidance.
Animals subject to predation pressure may
commonly evolve avoidance behaviors in
response to environmental conditions that
indicate increased risk of predation. Despite a
sacrifice in resource capture from foraging,
long-term selective pressures can reward
predator detection and avoidance.
ACKNOWLEDGEMENTS
We would like to thank Hannah ter Hofstede for
teaching us everything we know about bats and
trusting us with her valuable recording equipment.
We would also like to thank Don Carlos for
hanging out with us “mujeres nocturnas.”
AUTHOR CONTRIBUTIONS
All authors contributed equally.
ences. Neuroscience & Biobehavioral Reviews
25: 577-586.
Kotler, B.P. 1984. Effects of illumination on the
rate of resource harvesting in a community of
desert rodents. The American Midland
Naturalist 111: 383-389.
Lima, S.L., and L.M. Dill. 1990. Behavioral
decisions made under the risk of predation: a
review and prospectus. Canadian Journal of
Zoology 68: 619-640.
Dartmouth Studies in Tropical Ecology 2014
Pyke, G.H., H. R. Pulliam, and E.L. Charnov.
1977. Optimal foraging: a selective review of
theory and tests. Quarterly Review of Biology
52: 137-154.
Quinn, G.P., and M.J. Keough. 2002.
Experimental Design and Data Analysis for
Biologists. Cambridge, UK: Cambridge UP.
Rydell, J., Entwistle, A., and Racey, P.A. Timing
of foraging flights of three species of bats in
relation to insect activity and predation risk.
Oikos 1996: 243-252.
Shen, Y., Liang, L., Zhu, Z., Zhou, W., Irwin,
D.M., Zhang, Y., and D.M. Hillis. 2010.
Adaptive evolution of energy metabolism
genes and the origin of flight in bats.
Proceedings of the National Academy of
Sciences 107: 8666-8671.
Stiles, G., and A.F. Skutch. 1989. A Guide to the
Birds of Costa Rica. Comstock Publishing
Associates, Ithaca, NY.
Voigt, C.C., Sorgel, K., and D.K.N. Dechmann.
2010. Refueling while flying: Foraging bats
combust food rapidly and directly to power
flight. Ecology 91: 2908-2917.
55
Cuericí
SPATIAL PATTERNING IN THE ABUNDANCE OF FLORAL MITES PHORETIC ON
HUMMINGBIRDS
LARS-OLAF HӦGER AND ZACHARY T. WOOD
Faculty Editor: Matthew P. Ayres
Abstract: When resources are patchy and dispersal is limited, population stability in a landscape can become
sensitive to dispersal and extinction events in the metapopulation. Spatial patterns in abundance can reveal the extent
of patchiness and therefore the potential for metapopulation dynamics. We investigated how the distribution of
flowers across a range of scales affected the abundance of mesostigmatid mites that feed in hummingbird-pollinated
flowers and are phoretic on the hummingbirds. Abundance was relatively uniform among flowers within
inflorescences and negatively associated with the occurrence of other inflorescences within one meter. However,
higher numbers of inflorescences within five meters were associated with higher abundance of mites per flower,
perhaps because mite populations are patchy and exhibit localized aggregations. Organisms with relatively limited
dispersal capabilities depend highly on patches in close proximity for colonization and therefore tend to exist in
fragmented, spatially patchy metapopulations.
Key words: Bomerea alstroemeriaceae, patch structure, phoresy, metapopulations, mites
INTRODUCTION
Populations occur in space across a landscape.
Dispersal limitations (e.g., geographic barriers) or
physiological constraints (e.g., survival and
reproduction limits) can sometimes confine
subpopulations to patches within landscapes.
When organisms can disperse among patches,
even infrequently, the population of patches or
metapopulation can be stable, even though not all
patches are occupied and occupancy of some is
intermittent. As patches become distant beyond
reasonable dispersal, a tipping point may be
reached in which a metapopulation collapses and
most patches become unoccupied (Hanski, 1998).
Therefore, the stability of populations that exist as
metapopulations can be highly sensitive to the
scale of patches relative to the dispersal capability
of individuals.
Organisms that rely on patchily distributed
resources are obvious candidates for being
structured as metapopulations. We studied
phoretic mesostigmatid mites that live within
hummingbird-pollinated flowers of Bomerea
alstroemeriaceae at the Cuerici Biological Station
in Costa Rica. B. alstroemeriaceae flowers occur
in large inflorescences, 20 or more flowers on the
same stem. The mites move among plants in the
nares of feeding hummingbirds (Colwell and
Naeem, 1994). Inflorescences can be many meters
away from the nearest other inflorescence.
56
The patchy dispersion of B. alstroemeriaceae
inflorescences and the dependence of mites on
facilitated transportation suggest that the
dispersion of mites should be even patchier than
the floral resources that support them. We sought
to understand how spatial arrangements of B.
alstroemeriaceae affected mite abundance. We
tested three hypotheses: (1) mites are patchy
(aggregated) at the scale of individual flowers or
inflorescences due to the limitations of mite
movement, (2) mites are patchy at the coarser
scale of groups of inflorescences, or (3)
hummingbird facilitated dispersal is so efficient
that mite abundance is relatively uniform even at
coarse scales, implying that they are not subject to
metapopulation dynamics.
METHODS
We surveyed inflorescences of B.
alstroemeriaceae along a research trail at the
Cuereci Biological Field Station at approximately
2,600 meters above sea level on Cerro de la
Muerte, Costa Rica. We haphazardly selected
inflorescences and counted the number of flowers
in each inflorescence. We randomly sampled two
flowers per selected inflorescence and counted the
number of mites in each. We recorded the number
of additional B. alstroemeriaceae inflorescences
located within a 1, 5, and 10 meter radius of the
Dartmouth Studies in Tropical Ecology 2014
focal inflorescence (for data distributions, see
Figure 1).
We employed a nested, random effects
analysis of variance (ANOVA) to estimate how
much of the variation in mite abundance per
flower could be explained by the inflorescence to
which it belonged. That is, did flowers within an
inflorescence tend to have similar numbers of
mites?
We tested for the effects of nearby flowers and
inflorescences on the number of mites per flower
using a generalized linear model with a Poisson
distributed link function. We examined several
models with number of flowers per focal
inflorescence and number of additional
inflorescences within 0-1, 1-5, and 5-10 meters as
potential independent variables. We used AICc,
Log-likelihoods, and corresponding model weight
to compare the five competing models (Anderson,
2008). We used model averaging based on our
model weights to arrive at a final predictive model
(Anderson, 2008).
To evaluate whether there was clumping of
mite abundance at a coarser scale than 10 m, we
created and examined a map of mites per flower
throughout or study area.
RESULTS
Mite abundance was strongly structured by
individual inflorescences: 70% of the variation in
per-flower mite abundance was explained by the
inflorescence to which a flower belonged, and
only 30% of the variance was attributable to the
flowers within inflorescences.
In evaluating the predictors of mite abundance in
inflorescences, we found that two models were
substantially better than the other three competing
models (Table 1; ΔAICc= 54 between Models 1
and 2, and Model 3). These top two models had
identical AICc values and model weights (760 and
0.5, respectively); one model containing all the
predictors had higher goodness of fit (lower loglikelihood) but had two more parameters than the
alternative plausible model which only included
inflorescences between meters and 0-5 meters.
Because both models were equally plausible we
adopted a model-averaging approach (Anderson,
2008) to yield weighted parameter estimates for a
single predictive model of mite abundance (Table
1): 𝑁𝑁𝑁𝑁𝑁𝑁 = 6.67−0.15𝑁𝑁 − 1.74𝑁0−1𝑁 +
0.61𝑁1−5𝑁 − 0.11𝑁5 −10𝑁
In this model, Fi is the number of flowers per
inflorescence, and Ix-ym is the number of
inflorescences between x and y meters from the
focal flower. Inflorescences with more flowers had
fewer mites (Figure 2A). Within a 1 meter radius
the number of mites per flower decreased with the
number of inflorescences (Figure 2B). Within a 15 meter band the number of mites per flower
increased with number of inflorescences (Figure
2C). Between 5 and 10 meters, increased
inflorescence density decreased mite abundance in
the focal flower (Figure 2D). Overall, there was a
positive trend in mites per flower with number of
inflorescences within 10 meter (chi-square =
11.62, df=1, p=0.0006, Figure 3). We saw no
obvious large-scale trends in mite abundance
(Figure 4).
Table 1. Models predicting number of mites per flower.
Model #
Parameter estimate ± standard error
Intercept
Flowers per
inflorescence
Inflorescences
between 0 &
1m
Inflorescences
between 1 &
5m
Inflorescences
between 5 &
10m
AICc
1
6.92 ± 0.70
-0.03 ± 0.01
-1.52 ± 0.32
0.65 ± 0.14
-0.21 ± 0.10
2
6.42 ± 0.65
-1.96 ± 0.26
0.60 ± 0.09
3
7.95 ± 0.69
-1.88 ± 0.29
4
2.75 ± 0.28
5
4.28 ± 0.23
0.54 ± 0.09
Δ
LogL
Wi
0
374.65
0.5
0
377.21
0.5
54
405.13
0
46
401.28
0
91
424.4
0
57
Cuericí
58
Dartmouth Studies in Tropical Ecology 2014
DISCUSSION
Our results indicate that these mesostigmatid mites
exist in patchy subpopulations. Our finding that
flowers within the same inflorescence have similar
numbers of mites suggests that mites are able to
move readily about an inflorescence, possibly
without hummingbird facilitation. The result that
an increased number of flowers per inflorescence
or inflorescences within one meter decreases the
abundance of mites in each flower suggests that
mites are dispersing and spreading out within a
one meter patch. As this range seems beyond that
of an individual mite’s normal dispersal ability,
the observed spreading out is probably
hummingbird facilitated. Collectively, this implies
the existence of meaningful patches for the mites
at a scale of about a one meter radius.
The pattern changed between a one and five
meter radius. Mite abundance per flower increased
with inflorescence abundance between one and
five meters from the focal flower. This positive
effect on mite abundance suggests a population
subsidization effect from nearby patches. Having
more neighbor inflorescences increases the
frequency of colonization and immigration events.
This could be due to the relatively local feeding
patterns of hummingbirds.
The small negative effect on mite abundance
per flower from inflorescences outside a five
meter radius could be due to a different feature of
hummingbird behavior. Hummingbirds may pick
areas with higher inflorescence densities to feed,
so having a high density of inflorescences far from
the focal inflorescence could create a more
preferred patch for hummingbirds, which would
therefore decrease the frequency of visits to the
59
Cuericí
focal patch, decreasing the chance of patch
colonization. This remains consistent with the
interpretation that most colonization and
immigration events in our mite metapopulation are
among flowers within five meters. If so, the
maintenance of the mite metapopulation would
require other flowers within a relatively small
distance. Flowers within an inflorescence of B.
alstroemeriaceae open simultaneously and appear
to be relatively short-lived, so frequent
colonization events must be necessary to maintain
an extant metapopulation. Without regular
movement, the sub-populations of mites would
likely perish with their inflorescences. Thus our
study population of mesostigmatid mites could be
quite sensitive to landscape alterations that
increase the nearest neighbor distances of B.
alstroemeriaceae inflorescences to greater than
five meters.
Our study demonstrates the potential for
populations to be structured at multiple scales.
Hierarchies of demographic spatial structure might
be a common feature of populations that have
multiple modes of dispersal (e.g., mites that walk
among flowers vs. catching rides among
inflorescences). Understanding the scale of
patches in spatially heterogeneous populations can
be important to understanding population
dynamics and stability.
LITERATURE CITED
Anderson, D.R. 2008. Model based inference in
the life sciences: a primer on evidence (New
York; London: Springer).
Colwell, R.K., and Naeem, S. 1994. Life-History
Patterns of Hummingbird Flower Mites in
Relation to Host Phenology and Morphology.
In Mites, M.A. Houck, ed. (Springer US), pp.
23–44.
Hanski, I. 1998. Metapopulation dynamics.
Nature 396, 41–49
60
ACKNOWLEDGEMENTS
We would like to thank Matthew Ayres, Jessica
Trout-Haney and Vivek Venkataraman for their
help and guidance in this project.
AUTHOR CONTRIBUTIONS
Both authors contributed equally
Dartmouth Studies in Tropical Ecology 2014
FEMALE MATE CHOICE AND MALE LEKKING BEHAVIOR IN DROSOPHILA
AMELIA L. RITGER
Faculty Editor: Matthew P. Ayres
Abstract: The effects of female mate choice on males can be displayed in three major ways: ornamental displays,
resource-based mating systems, and non-resource-based mating systems such as leks. Drosophila fruit flies are
among the organisms where female mate choice, and its consequences, have been demonstrated in the laboratory.
Field studies of sexual selection in Drosophila are rare by comparison. Drosophila males in the Cuericí Biological
Reserve in Costa Rica were found apparently competing for females at a mushroom patch. Given that Drosophila
larvae consume decaying fruit and fungi, I hypothesized that the Drosophila mating system was resource-based and
that fly preferences within the patch would be structured by mushroom size. Results were contrary in that there was
no relationship between mushroom size and the intensity of male competition or the number of female visits to
mushrooms. However, Drosophila were attracted to some mushrooms more than others and male competition was
highest on the same mushrooms that females were most likely to visit. Thus, the Drosophila were using the
mushrooms as leks. Additionally, dominant male flies were much larger than non-dominant male flies, suggesting
that larger males had higher mating success than smaller males.
Keywords: Drosophila, lekking, mating systems, female mate choice
INTRODUCTION
Female mate choice drives sexual selection on
males across many taxa. Females may choose
males by their physical attributes (Brown 1978),
the quality of territory they control (Pleszczynska
1978), or their behavior (Janetos 1980). One
common sexual selection model suggests that
female mate choice drives ornamental displays in
males. For example, male peacocks often have
exaggerated ornamental displays preferred by
females even though they come at the cost of
higher predation risk (Andersson 2008). Mate
choice can also create resource-based mating
systems, which exert pressure on males to defend
a territory that provides the female with a
resource. For example, male Mediterranean pollen
wasps defend water collection sites and flowers,
both of which are important for female brood care
(Groddeck et al 2004). In non-resource-based
mating systems such as leks, males display in an
area lacking a resource where females aggregate
solely for the purpose of mate selection,
presumably because male success in leks tends to
indicate good genes. For example, lekking male
cichlids dig and defend spawning pits where
females congregate and mate with the defending
male (Nelson 1995).
Many Drosophila species breed in decaying
fruit or fungi (Borror et al 1989) and are
frequently used for laboratory studies on genetics
and evolution due to their short generation time
(10-30 days) and relatively short adult life span
(Spieth 1974). Female Drosophila have been
found to discriminate between males based on
mutant type (Rendel 1945), ability to successfully
defend food resources (Hoffmann 1987), and
lekking display size (Aspi and Hoffmann 1998).
In the Cuericí Biological Reserve, Costa Rica,
I encountered aggregations of Drosophila males
aggressively fighting each other in a mushroom
patch. There appeared to be a dominant male on
each mushroom and a number of floater males and
females flying around the mushrooms. I observed
females exploring the undersides of the
mushrooms as though they were searching for
oviposition sites, implying that the mushrooms
were a larval feeding resource. I tested the
hypothesis that this Drosophila mating system was
resource-based and that female choice would be
based on mushroom size. If the mushrooms were
being used as a resource for feeding larvae, then it
would be logical that both sexes would prefer
larger mushrooms. If the Drosophila were not
using the mushrooms as a resource but merely as a
lek, there could be some mushrooms that were
more favored than others but this would not
necessarily be related to mushroom size. I also
tested the theoretical expectation that male
dominance would be related to body size (Tokarz
1985).
61
Cuericí
METHODS
I studied the Drosophila aggregated at a patch of
unidentified fruiting basidiomycetes (Fig. 1) along
the Cuericí canyon trail approximately 50 meters
before the stream on the uphill slope. Observations
were made between 10:00 and 15:00 on 30-31
January 2014. I used a compound microscope at
10-50x magnification and identified the flies as
Drosophila by examining their body size
(typically 2-4 mm), abdominal and sternoplural
bristles, and wing venation (Borror et al 1989). I
identified the dominant male as the only male on
each mushroom that was aggressive towards and
won fights against floater males that occasionally
landed. I identified floater males as the males
without control over a mushroom or as
competitors who lost a fight against the dominant
male. I measured ambient air temperature around
the mushroom patch every 5 minutes using a
Hoboware Data Logger JKST thermocouple.
Observational Study
I assigned numbers to 43 haphazardly selected
mushrooms in the patch (Fig. 1) and recorded
observations of mushrooms with dominant males
for 28 five-minute periods. During each trial
period, I counted the number of male competitors
that landed on the focal mushroom, whether or not
the dominant male was replaced by a male
competitor and the number of females that landed
62
on the mushroom. I also noted whether or not the
dominant male mated with a female and looked for
the presence of eggs underneath and inside the
mushroom. I collected 16 dominant males and 21
floater males and measured body length, body
width, and thorax length. After the trial period, I
also measured the length, width, and height of
each mushroom.
Mushroom Desirability Experiment
To test how desirable each mushroom was for
males, I removed (by aspiration) the dominant
male on each mushroom and measured the amount
of time it took for a new dominant male to take
over the mushroom. I removed the new dominant
male and continued removing new dominant males
for one minute, counting the total number of new
dominant males that landed on each mushroom.
Male Dominance Experiment
To test if the dominant males were able to reestablish their dominance after being temporarily
removed from their mushroom, I aspirated 10
dominant males and 10 floater males and
respectively dyed them with orange or green
fluorescent dye. I then waited 24 hours and
counted the number of previously dominant males
that were dominant on mushrooms and the number
of floater males that became dominant on
mushrooms.
Statistical Analyses
To analyze if the mushrooms were being used as a
resource and if male and female flies were
attracted to the same mushroom, I regressed the
total number of females against the number of
male competitors, the number of females against
mushroom size, and the number of male
competitors against mushroom size. I incorporated
body length, body width, and abdomen length into
a single metric of body size using a principal
components analysis (PCA). The first principal
component axis explained 80% of the variation
with loadings of 0.903, 0.899, and 0.877 for body
length, body width, and abdomen length,
respectively. To examine the relationship between
male dominance and body size, I compared the
first principal component axis with fly social
status using a two-tailed t-test (Jolliffe 2002). All
analyses were performed using JMP® 10
statistical software (SAS Institute, Cary, NC).
Dartmouth Studies in Tropical Ecology 2014
RESULTS
Drosophila were only active at the mushroom
patch between 10:00 and 15:00 on both
observation days. Shaded ambient temperatures
within those time periods were 10.8-11.8 °C on 30
January and 10.8-13.9 °C on 31 January.
mushrooms with dominant males.
Observational Study
Mushroom size was not related to the number of
females (Fig. 2A) nor to the number of male
competitors (Fig. 2B). However, the total number
of female visits and the number of male
competitors were positively correlated (Fig. 3).
The smallest dominant male was larger than the
largest floater male (Fig. 4). During the course of
the 28 five-minute periods, I observed one mating
by a dominant male and 4 cases where the
dominant male was replaced during the trial. No
eggs were found under or inside any of the
Mushroom Desirability Experiment
There was no relationship between mushroom size
and the time until arrival of a new dominant male
on the mushroom (P = 0.63, r2 = 0.01). There was
also no relationship between time of arrival of a
new dominant male and number of male
competitors (P = 0.66, N =28) or number of
females (P = 0.25, N = 28).
Male Dominance Experiment
18 hours after experimental dusting with
fluorescent powder, only one orange dyed
(previously dominant) male was found in the area.
It was not on a mushroom and appeared moribund.
63
Cuericí
DISCUSSION
Neither female nor male flies showed any
discrimination based on mushroom size (Fig. 2),
yet both sexes were attracted to the same
mushrooms (Fig. 3). Although Drosophila larvae
have been known to consume mushrooms during
development, I did not find any eggs present under
the mushroom. However, I did see a mating occur
on one of the mushrooms. Evidently, the
mushrooms served as a mating territory but did not
provide the flies with food. That is, the Drosophila
males were lekking on the mushrooms and the
females were aggregating to select mates based on
the results of dominance contests among males.
My observations suggested that there might be
a “home-field” advantage within this lek system
because the dominant fly tended to remain
dominant even if the competing male was larger.
This could be because ownership status compels
territorial residents to value a contested area more
than intruders, and thus residents have the
potential to gain more from winning and/or risk
more by defeat (Enquist and Leimar 1987). Future
studies could test for a home field advantage by
experimental removal and later release of
dominant males.
The strong relationship between body size and
male social status suggests intense selection for
large body size in Drosophila, especially because
the one observed mating was by a dominant male
(Fig. 4). The results of the male dominance
experiment would have been informative if there
had been re-sightings the next day. I think that the
flies were excessively dusted with the fluorescent
powder and died as a result. However, it could
also be that the typical male fly only competes in
the lek for a single day. This would be consistent
with the limited seasonal variation in this tropical
environment and the typically short generation
time of Drosophila. Additional marking studies
with less aggressive dusting could indicate
whether or not flies are able to retain status as
dominant male over successive days.
Mating systems such as leks arise when
female mate choice places strong pressure on
males to compete or make themselves more
attractive to the females by providing direct
benefits (such as a high quality territory) or
indirect benefits (such as good genes). Studying
lekking behavior and other behaviors by males in
response to female mate choice allows us to better
understand the evolutionary basis for intersexual
selection.
LITERATURE CITED
Andersson, M. 2008. Sexual selection, natural
selection, and quality advertisement.
Biological Journal of the Linnean Society 17:
375-393.
Aspi, J., and Hoffmann, A. 1998. Female
encounter rates and fighting costs of males are
associated with lek size in Drosophila
mycetophaga. Behavioral Ecology and
Sociobiology 42: 163-169.
Bateman, A.J. 1948. Intra-sexual selection in
Drosophila. Heredity 2: 349-368.
Borror, D.J., and White, R.E. 1970. A Field
Guide to Insects: America North of Mexico.
Houghton Mifflin Co., Boston, MA.
Borror, D.J., Triplehorn, C.A., and Johnson, N.F.
1989. An introduction to the study of insects,
6th edition. Saunders College Publishing,
Philadelphia, PA.
Brown L.P. 1978. Polygamy, female choice, and
the mottled sculpin, Cottus bairdi. PhD
dissertation, Ohio State University.
Enquist, M., and Leimar, O. 1987. Evolution of
fighting behaviour: the effect of variation in
resource value. Journal of Theoretical Biology
127: 187-205.
Groddeck, J., Mauss, V., and Reinhold, K. 2004.
The resource-based mating system of the
Mediterranean Pollen Wasp Ceramius
fonscolombei Latreille 1810 (Hymenoptera,
Vespidae, Masarinae). Journal of Insect
Behavior 17: 397-418.
Hoffmann, A. 1987. A laboratory study of male
territoriality in the sibling species Drosophila
melanogaster and D. simulans. Animal
Behaviour 35: 807-818.
Janetos, A. 1980. Strategies of female mate
choice: a theoretical analysis. Behavioral
64
ACKNOWLEDGEMENTS
I thank Matthew Ayres for assisting me with the
principal components analysis. Thanks also to
Annie Fagan, Emily Goodwin, and Aaron
Mondshine for peer-reviewing my paper and
Jessica Trout-Haney and Vivek Venkataraman for
reviewing my paper.
Dartmouth Studies in Tropical Ecology 2014
Ecology and Sociobiology 7: 107-112.
Jolliffe, I.T. 2002. Graphical representation of data
using principal components. Principal
Components Analysis, 78-110. Springer, New
York, NY.
Nelson, C. 1995. Male size, spawning pit size, and
female mate choice in a lekking cichlid fish.
Animal Behaviour 50: 1587-1599.
Pleszczynska, W.K. 1978. Microgeographic
prediction of polygyny in the lark bunting.
Science 201: 935–937.
Rendel, J.M. 1945. Genetics and cytology of
Drosophila subobscura. Journal of Genetics
46: 287-302.
Spieth, H. 1974. Courtship behavior in
Drosophila. Annual Review of Entomoloy 19:
385-405.
Taylor, C.E., and Kekic, V. 1988. Sexual selection
in a natural population of Drosophila
melanogaster. Evolution 42: 197-199.
Tokarz, R. 1985. Body size as a factor determining
dominance in staged agonistic encounters
between male brown anoles (Anolis sangrei).
Animal Behaviour 33: 746-775.
65
Cuerici
SIZE DISTRIBUTION AND AGGRESSION IN RAINBOW TROUT, ONCORHYNCHUS
MYKISS
ALLEGRA M. CONDIOTTE, NATHANAEL J. FRIDAY, AND ADAM N. SCHNEIDER
Faculty Editor: Matthew P. Ayres
Abstract: Many organisms use aggression to increase access to limited resources. However, aggressive behaviors are
energetically costly, and carry a risk of injury that could reduce an individual’s fitness. These risks are higher when
aggression is among similarly-sized individuals, which may be capable of inflicting equal damage. Thus, the
distribution of body sizes could have an effect on the frequency of aggressive behaviors. We tested whether
agression in aquacultured Oncorhynchus mykiss (rainbow trout) varied with the size distribution of interacting fish.
We surveyed O. mykiss in two enclosures, one with mostly large trout and some small trout and another with mostly
small trout and some large trout. We counted the number of fin tears resulting from bites. The trout showed more
signs of aggression in the smaller-size population than in the larger population, but neither sex nor fish length was
related to frequency of fin tears. Larger fish in the small-size population had more tears on their dorsal, anal and
caudal fins, suggesting that the majority of attacks on larger competitors was from behind, which is a significantly
less risky approach, while smaller fish in this population had more tears in their pectoral and pelvic fins, suggesting
that they are typically attacked from above or ahead, a strategy that is riskier for the attacker but may yield greater
damage. The size distribution of interacting individuals can influence aggressive behaviors among organisms that
live in groups.
Key words: aggression, intraspecific competition, size distribution, rainbow trout, Oncorhynchus mykiss
INTRODUCTION
Individuals within a population compete for access
to limited resources, from prey items to mating
opportunities. Organisms can sometimes increase
their success in intraspecific competition by
asserting dominance over other conspecifics
through aggressive behaviors. Aggression can
increase an individual’s access to a limited
resource, but it comes at a cost through loss of
feeding opportunities and the risk of damage
(Jaeger 1981).
The extent to which aggression is favored or
disfavored is likely to be influenced by the social
context. Animals can limit injury risks by tending
to direct aggression toward smaller, less
threatening individuals. In this case, larger animals
should be more aggressive on average because
they have more safe targets. On the other hand,
animals of similar size are more likely to compete
for the same set of resources; therefore, aggression
against individuals of the same size yields greater
rewards than against smaller individuals.
We examined the influence of differing size
distributions of rainbow trout (Oncorhynchus
mykiss) on the extent and nature of conspecific
aggression. Specifically, we compared the
aggression preferences of two groups of O. mykiss,
66
one with mostly large fish but some small fish and
one with mostly small fish but some large fish, by
examining the number of fin tears resulting from
bites by conspecifics.
If fish direct their aggression to maximize the
likelihood of successfully exerting dominance
while simultaneously minimizing the risk of
injury, there should be an inverse relationship
between fish size and fin damage (more fin tears
in small fish). However, if fish direct their
aggression primarily to maximize competitive
gain, fin damage should tend to be greatest in the
most abundant size class.
METHODS
Data Collection
We surveyed the length of fish in two enclosures
in a small-scale trout farm in Cuerici, Costa Rica
on 29 and 30 January 2014. One enclosure
contained trout primarily of large body sizes
(Interquartile Range = 37 cm to 44 cm), while the
second enclosure contained trout of mostly small
body sizes (IQR = 25 cm to 32.5 cm). Population
densities were not known precisely between
enclosures, but it seemed evident that density was
much higher in the small fish enclosure. A mesh
Dartmouth Studies in Tropical Ecology 2014
barrier separated the enclosures. Fish had been in
the present enclosures for about six months.
In each enclosure, we attracted trout with food
and captured them with a net. We put the captured
trout in a submerged holding cage, then transferred
them one at a time to a measuring tank, where we
assessed them for fin damage. We counted the
number of tears between the spines on each fin
(Figure 1). We did not include bite marks and
severed spines, which scar and usually do not
regrow. We used fin tears as a measure because
they heal relatively quickly and are thus an
indicator of recent aggression, as opposed to other
wounds that could have been inflicted much
earlier, including when the fish were in different
enclosures.
We also recorded the length and sex (based on
jaw structure) of each trout before tagging and
releasing it. We did not include fish under 20 cm
in length because they could not be sexed.
Statistical Analysis
We used a Kolmogorov-Smirnov test to compare
the distributions of sizes in the two enclosures. We
compared the number and distribution of fin tears
between the two enclosure treatments. We
employed a general linear model that included
enclosure, sex, length, enclosure x sex, and
enclosure x length as predictors of total number of
fin tears. We used the same model to evaluate
patterns in the first and second axes from a
principal components analysis of number of tears
on each of the seven individual fins (Table 1).
RESULTS
The variance was very similar between enclosures
(SD = 6 vs. 7; F24,49=1.36), so the coefficient of
variation was greater in the enclosure with smaller
fish (CV = 100 x SD / mean = 23.0% vs. 14.0%).
The Kolmogorov-Smirnov test showed clear
differences between the enclosures in the size
distributions of fish (P<0.0001). This reflected
differences in mean body lengths and a tendency
for the enclosure with large fish to be skewed to
the left while the enclosure with smaller fish was
skewed to the right (Figure 2).
Enclosure was the only factor that had a
significant effect on total number of fin tears, with
more tears in the enclsoure with smaller fish (F1,69
= 5.48, P = 0.022; Table 1). The first axis of the
PCA was interpretable in about the same way as
total tears (all positive loadings; Table 2). The
second axis captured variation in the distribution
of damage (high scores reflect more tears in
pectoral fins and fewer in dorsal fins). The
interaction between length and treatment was
significantly related to the second principal
component (F1,69 = 6.48, P = 0.013; Table 1). In
the enclosure with small fish, but not the enclosure
with large fish, PC-2 scores decreased with fish
Table 2: Loading values for the first two principal
components of tears per fin.
Source Fin
Dorsal
Pectoral Left
Pectoral Right
Anal
Caudal
Pelvic Left
Pelvic Right
Percent Variation Explained
PC1
0.42
0.23
0.10
0.42
0.44
0.41
0.47
25.2
PC2
-0.44
0.35
0.68
-0.29
-0.13
0.24
0.25
18.8
67
Cuerici
size (Figure 3); this indicated that relatively small
fish in this enclosure tended to show more damage
to the dorsal fins, while relatively larger fish
showed more damage to the pectoral fins.
Table 1: Results of tests for enclosure, sex, and length on the number and distribution of fin tears.
Source
Enclsoure
Sex
Sex*Enclosure
Length
Length*Enclosure
Error
Total tears
F Ratio
p
5.48
0.022
1.74
0.19
0.18
0.68
0.79
0.38
0.62
0.43
DF
1
1
1
1
1
69
Principal component 1
F Ratio
P
3.97
0.05
2.38
0.13
0.24
0.62
2.14
0.15
0.22
0.64
Principal component 2
F Ratio
p
1.23
0.27
2.20
0.14
0.03
0.87
1.84
0.18
6.48
0.013
A. Large
7
Number of Individuals
6
5
4
3
2
1
20
22.5
25
27.5
30
32.5
35
37.5
40
42.5
45
47.5
50
0
Size Class (cm)
Figure 2: Distribution of trout size classes in the (A) large-size enclosure (mean ± SD = 41 ± 6 cm), and (B)
small-size enclosure (mean ± SD = 29 ± 7 cm). The small-size enclosure also included some fish under 20
cm in length, which were not measured.
A. Large
14
12
Fin Tears
10
8
6
4
2
0
25
35
45
55
Body Length (cm)
Figure 3: Number of fin tears in the (A) large-size enclosure and (B) small-size enclosure. Dashed lines
show the non-significant lines of best fit. Note that axes are scaled differently.
68
Dartmouth Studies in Tropical Ecology 2014
A. Large
3
Distribution of Fin Tears
2
1
0
25
35
45
55
-1
-2
-3
Body Length (cm)
Figure 4: The second principal component of fin tears plotted against body length in the (A) large-size
enclosure, and (B) small-size enclosure.
DISCUSSION
Intraspecific aggression among fish could
theoretically be driven by either fish
maximizing success and minimizing injury,
which would involve mostly attacking smaller
fish, or maximizing competitive gain by
attacking the competitors closest in size. In
our study, Oncorhynchus mykiss showed more
signs of aggression in the population with
smaller fish that were relatively more varied
than in the population with larger fish that
tended to be closer in relative size.
Neither sex nor individual-size was related
to trout aggression, and the size of the most
damaged fish did not differ between the largesize and small-size fish populations.This
suggests that aggression was undirected
among larger fish, rather than being
strategically directed to minimize risk or
maximize reward, while in the population of
smaller fish, where relative size variation was
greater, aggression was more evident (more
tears in general) and the pattern of tears was
structured by size.
Larger fish in the small-size population
had relatively fewer tears on their pectoral and
pelv fins and more tears on their dorsal, anal
and caudal fins, suggesting that the majority of
attacks on larger competitors were from
behind, which is a significantly less risky
approach. In contrast, smaller fish in this
population had more tears in their pectoral and
pelvic fins, suggesting that they are typically
attacked from above or ahead, a strategy that
is riskier for the attacker but may yield greater
damage. This implies that length, or relative
length within a population, may have an effect
on aggressive strategy, perhaps based on
reprisal avoidance.
Our study was unable to separate effects
that might have been due to fish density. The
trout population with smaller fish had more
fish per cubic meter than the other enclosure
(Carlos Solano, personal communication), but
we were unable to estimate relative biomass
per cubic meter. High population density can
cause higher levels of competition and
increase encounter rates, which can lead to
greater frequency of aggressive behaviors
(Skogland 1983). However prior studies at
Cuerici Biological Station have shown that
population density had no effect on fish body
condition (Costello et al. 2012). This is
consistent with the increased aggression in the
small size population being more related to the
size distribution than to the population density.
However, there would be value in further
studies that more effectively isolate effects of
size distributions vs. density.
Once an aggression event occurs, the
presence of a weaker, injured fish may incite
other trout to become aggressive against each
69
Cuerici
other. In a population skewed toward smaller
fish, where the few large trout may be more
likely to attack and seriously injure smaller
competitors, this could lead to a higher level
of aggression overall.
Our results indicate that the size
distribution of a population may affect the
distribution of damage on a fish’s body. A
possible explanation for the damage on
smaller fish being located more on the front
and top of the body relative to larger indivials
is because larger fish are likely to be more
willing to risk attacking a fish from the front.
Future studies could examine this pattern
further. Also, focused observational studies of
trout aggression might shed light on the costs
and benefits of aggression vs. non-aggression.
When resources are limited, many
organisms direct aggression toward
conspecifics to expand their access to
resources and presumably increase their
fitness. However, the costs of aggressive
behaviors, including the risk of injury, may
put limits on the extent of aggression. Our
results indicate that size distribution of
interacting individuals can influence
aggressive behaviors, and that aggression can
change with age, size, or position in group
social hierarchy.
LITERATURE CITED
Costello, R.A., Deffebach, A.L, Gamble,
M.M., and Kreher, M.K. 2012. Effects of
density on farmed Oncorhynchus mykiss
body condition and implications for
economic sustainability of small-scale
fisheries. Dartmouth Studies in Tropical
Ecology 22: 62-68.
Skogland, T. 1983.The effects of density
dependent resource limitation on size of
wild reindeer. Oecologia 60:156-168.
Jaeger, R.G. 1981. Dear enemy recognition
and the costs of aggression between
salamanders. The American Naturalist
117: 962-974.
70
ACKNOWLEDGEMENTS
We thank Carlos Solano for the use of his
aquaculture pools and his assistance in data
collection. We also thank Jessica Trout-Haney
and Vivek Venkataraman for their advice on
the writing process, and Matthew Ayres for
his direction.
AUTHOR CONTRIBUTIONS
All authors contributed equally in the data
collection and writing process.
Dartmouth Studies in Tropical Ecology 2014
AQUATIC COMMUNITY COMPOSITION OF THE RÍO CLARO ESTUARY
ALLEGRA M. CONDIOTTE, NATHANAEL J. FRIDAY, LARS-OLAF HÖGER,
AARON J. MONDSHINE, ELLEN R. PLANE, AND AMELIA L. RITGER
Faculty Editor: Matthew P. Ayres
Abstract: Species distributions are determined by environmental factors, which may create abrupt or gradual
ecotones between habitats. Estuaries mark a transition zone between freshwater and marine systems, and create a
unique overlap of conditions that change temporally with tidal flow. Ecotones like estuaries create edge habitat that
may increase species richness by providing additional niches; however, temporal salinity shifts create an
osmoregulation challenge that may reduce species diversity. Species assemblages could change suddenly near the
edge of high tide incursions, or individualistic responses in tolerance could create a gradual gradient in community
structure along the estuary. In fact, we found a gradual transition in species composition from the mouth of the river
upstream, with an intermediate zone of high species richness at about 1100 meters that could not be explained by a
salinity gradient. Our results demonstrate the importance in defining niches of factors more diverse than salinity,
including substrate and flow rate.
Key words: community dynamics, ecotone, estuary, Río Claro
INTRODUCTION
Species distributions are limited by abiotic and
biotic factors that vary over geographic distances.
Such ecostones can be graded or abrupt. Resource
availability, tolerance for environmental
conditions, species interactions, and niche
partitioning all influence the degree of overlap of
species assemblages at ecotones. Estuaries form an
ecotone between freshwater and marine
ecosystems. Tidal flow alters these ecosystems on
a temporal scale as the high tide shifts marine and
brackish conditions up river. Changing water
depth and direction of flow can also alter the
riverbed substrate, moving cobbles and organic
particulates of varying sizes, affecting river
characteristics even far above the high tide line.
Estuary systems can contain generalist species
that are quite tolerant of changes in water
chemistry, as well as relatively specialized species
that are more restricted in where they occur. Many
species in the Río Claro spend part of their life
cycle in the Pacific Ocean (Lyons and Schneider
1990, Schneider and Lyons 1993), and some
marine species use estuaries as temporary nursery
sites (Claridge et al. 1986). The Osa Peninsula is
unusual in a way that might matter to the estuary
ecosystem because most freshwater fish species in
the area are derived from marine species
(Constantz et al. 1981).
We studied the Río Claro for estuary species
richness, species assemblages, and the nature of
shifts in the community structure. Ecotones such
as the estuary boundary between fresh and
saltwater provide a type of edge habitat that can
facilitate species richness by encouraging
colonization by edge-adapted species (Naiman et
al. 1988, Ward et al. 1999).
However, the temporal variation in water
chemistry in estuaries poses a physiological
challenge to aquatic species. Tidal swings in
osmolarity and pH, for example, might result in
reduced species richness at the ecotone. The
ecotone associated with tidal fluxes might produce
edges in the community structure as entire groups
of organisms meet a threshold, or species might
respond quite individualistically to the many
ecological factors correlated with tidal fluxes, in
which case there would be a gradual shift in
species assemblages along the estuary.
METHODS
Data Collection
We assessed species composition along the Río
Claro in Corcovado National Park, Costa Rica on
4-5 January, 2014. We sampled a site every 100
meters between 286 and 2050 meters upstream of
the river mouth (Fig. 1). At each site we recorded
water temperature, pH, and salinity using a YSI
Model 63 handheld system. We also noted light
condition, water clarity, presence of algae,
approximate water depth, substrate type, and water
flow.
71
Corcovado
Table 1. First and second principal components axis comparing community structure across the river.
Taxa
Coenobita sp.
Uca sp.
Neritina latissima
Cochliopina tryoniana
Atyidae
Bufo marinus
Actinopterygii morphotype 1
Sphoeroides annulatus
Actinopterygii morphotype 2
Agonostomus monticola
Actinopterygii morphotype 3
Unique Algal Morphotype
Apogon pacificus
Apogon dovii
Cryptoheros myrnae
Gobidae
Bathygobius soporator
Bathygobius ramosus
Lutjanus inermis
Lutjanus guttatus
Centropomus undecimalis
Mugil curema
Pseudophallus starksi
Percent variation:
Principal Component 1
0.17
0.07
-0.67
-0.45
-0.86
-0.39
0.37
0.68
0.38
-0.45
0.37
0.18
0.68
0.53
0.32
0.41
0.39
-0.22
0.32
0.62
0.38
0.18
-0.6
Principal Component 2
-0.54
-0.52
0.46
0.32
0.36
0.39
-0.71
0.48
-0.57
0.29
0.36
0.49
0.33
0.41
0.10
0.33
0.55
-0.04
0.10
0.39
0.29
-0.09
0.16
21.50%
16.00%
We spent 10 minutes at each site categorizing
aquatic vertebrate and invertebrate species and
assessing their relative abundances. When
possible, we recorded total counts of individuals of
a species. We estimated abundances of highdensity organisms using quadrats of approximately
1 m2. We classified fish species using the
FishBase.org database.
Statistical Analysis
We square root transformed the species richness
data and then employed a Principal Components
Analysis to reduce the dimensionality of
community structure (Table 1). We described pH
and salinity along the study transect with a 4parameter logistic function (Equation 1):
1
f (x) =
´ (Ymax -Ymin ) +Ymin
B´( xmid -x)
1+ e
Equation 1
where x is the measured value of salinity or pH, B
is the slope, xmid is the inflection point, and Ymax
72
and Ymin are the maximum and minimum for the
function.
Table 2. Estimated parameters for logistic fits of
salinity and pH as a function of distance from river
mouth (Equation 1).
Salinity
pH
B ± SE
-0.003 ± 0.02
-0.002 ± 0.003
Ymax ± SE
35.0 ppt
8.14 ± 0.17
Ymin ± SE
0.09 ± 0.02 ppt
7.78 ± 0.06
Xmid ± SE
1409 ± 335 meters
We employed a 3-parameter unimodal
function to describe species richness along the
transect (Equation 2):
f (x) = a ´ (dist - h)2 + k
Equation 2
Dartmouth Studies in Tropical Ecology 2014
where a is the concavity of the curve, dist is the
distance from the mouth of the river, h is the xvalue of the vertex, and k is the y-value of the
vertex. We constructed a floating bar chart to
examine the differences in range among species.
RESULTS
Salinity and pH both decreased with distance from
the mouth of the river (Table 2, Figs. 2-3). Each
aquatic species had a different spatial distribution
along the Río Claro (Fig. 4). Species richness
peaked (k = 7.3 ± 0.7 species) at an intermediate
distance of 1132 ± 98 meters from the mouth (Fig.
5).
The first axis from a PCA of community
structure indicated a gradual, approximately linear,
change in the community along the length of the
river (Table 1, Fig. 6). The second axis of
community structure indicated a relative rapid
change in structure in the lower river with relative
stasis above 800 meters. Patterns in PC-2 reflected
that two species of Actinoptergii fish were
abundant near the river mouth and rare higher in
the river (Table 1, Fig. 7).
Figure 1. The sites at which we measured stream characteristics and community structures. Waypoints are
overlaid on a GoogleTM Earth map of the first 2.5 kilometers of the Río Claro. ‘RioCMouth’ marks the defined
mouth of the river.
73
Corcovado
Figure 2. Water salinity (measured 0 to 2 hours after low
tide on 05 February 2014) decreased with distance from
the mouth of the river.
Figure 3. pH (measured 0 to 2 hours after low tide on 05
February 2014) decreased with distance from the mouth
of the river.
Figure 4. The observed range of aquatic species found along the Río Claro for each of 23 taxa. Species
such as Coenobita sp. and Actinopterygii morphotype 2 were found closer to the ocean whereas species
such as Pseudophallus starski and Neritina latissima were found furthest upstream.
74
Dartmouth Studies in Tropical Ecology 2014
Figure 5. Species richness peaked at 1132 ± 98 meters
from the mouth of the river, with 7.3 ± 0.7 species.
Figure 6. The first axis of community structure showed a
gradual and approximately linear transition in structure
along the river from the river mouth to 2000 meters
above the mouth (P<0.0001, r2 = 0.58).
DISCUSSION
Our results indicated a gradual transition in
species composition and abundance upstream from
the mouth of the river (Fig. 6, Fig. 7). The ecotone
did not create a congruent threshold for multiple
species. Apparently, the niches of our 23 study
taxa were defined by a diversity of different
environmental features (probably a combination of
biotic and abiotic factors). The result was that
species distributions were quite individualistic.
Species richness peaked at about 1100 meters
Figure 7. The second axis of community structure vs.
distance from the river mouth, P = 0.0043, r2 = 0.49
from the mouth of the river. This marked a
transition zone in which marine and freshwater
species overlapped.
Additionally there were four fish species that
occurred only in the transition zone. Evidently,
this area represents a unique set of conditions not
available at any other point along the stream
system. The low salinity (0.0-0.1 ppt) in this area
suggests that factors other than salinity, possibly
including mixed sediment size, varying water
flow/depth, and intermittent shading of the river
are responsible for the high community richness.
River width likely affects community
composition and species abundance as well. As
stream or river width changes, so does substrate
size, flow rate, and solar input. For example, the
flat snail Neritina latissima began to appear only
once cobble size and water flow increased, which
seemed to be where their body shape gave them an
advantage in clinging to large cobbles in a fastmoving water column. We also saw that average
fish size tended to decrease with distance upstream
for those fish species with broad ranges. This
suggests that life stage dictates a portion of
community structure in the Río Claro.
The principal components analysis suggested
the importance of certain species as drivers of
community composition. In sites on the upstream
side of 1100 meters, Pseudophallus starksi,
Neritina lattissima, and Atyidae sp. became
important elements of the community (Table. 1).
75
Corcovado
In sites downstream of 1100 meters, Sphoeroides
annulatus and Actinopterygii became more
important members of the community. This
change in species likely indicates changes in
trophic structure. The species becoming more
abundant upstream were frequently at low trophic
levels (e.g., algal consumers), while at least one of
the notable downstream species (Sphoeroides
annulatus) is a known predator at multiple trophic
levels (Chávez Sánchez et al., 2008). This may
reflect there being relatively greater energy
availability in communities near the ocean,
allowing for more species at higher trophic levels.
The shift in community structure and species
richness along the Río Claro lends insight into the
factors affecting transitional zones. The
LITERATURE CITED
Bell, J.D., Pollard, D.A., Burchmore, J.J.,
Pease, B.C., and Middleton, M.J. 1984.
Structure of a fish community in a temperate
tidal mangrove creek in Botany Bay, New
South Wales. Australian Journal of Marine
and Freshwater Research 35: 33-46.
Chávez Sánchez, M.C., Álvarez-Lajonchère,
L., Abdo de la Parra M.I., and García Aguilar,
N. 2008 Advances in the culture of the
Mexican bullseye puffer fish Sphoeroides
annulatus, Jenyns. Aquaculture Research. 39:
718–730.
Claridge, P.N., Potter, I.C., and Hardisty,
M.W. 1986. Seasonal changes in movements,
abundance, size composition and diversity of
the fish fauna of the Severn Estuary. Journal
of the Marine Biological Association of the
United Kingdom 1: 229-258.
Constantz, G.D., Bussing, W.A., Saul, W.G.
1981. Freshwater fishes of Corcovado
National Park, Costa Rica. Proceedings of the
Academy of Natural Sciences of Philadelphia
133: 15-19.
76
importance of factors other than salinity, including
substrate and flow rate, in determining community
composition, merits further study. These physical
changes in the environment can influence optimal
organism morphology, influence the range of
niches that are available, and thereby shape
community structure.
ACKNOWLEDGEMENTS
We thank Matthew Ayres, Jessica Trout-Haney,
and Vivek Venkataraman for their assistance.
AUTHOR CONTRIBUTIONS
All authors contributed equally in the data
collection and writing process.
Gore, P. 2000. Cluster Analysis. Pages 297321 in H.E.A. Tinsley and S.D. Brown,
editors. Handbook of Applied Multivariate
Statistics and Mathematical Modeling.
Academic Press, Waltham, MA.
Lyons, J. and Schneider, D.W. 1990. Factors
influencing fish distribution and community
structure in a small coastal river in
southwestern Costa Rica. Hydrobiologia 1: 114.
Martino, E.J., and Able, K.W. 2003. Fish
assemblages across the marine to low salinity
transition zone of a temperate estuary.
Estuarine, Coastal and Shelf Science 56: 969987.
Quinn, G., and Keough, M. 2002. Multiple
and complex regression. Pages 111-154.
Experimental Design and Data Analysis for
Biologists. Cambridge University Press,
Cambridge, UK.
Schneider, D.W., Lyons, J. 1993. Dynamics
of upstream migration of two species of
tropical freshwater snails. Journal ofthe North
American Benthological Society 121: 3-16.
Dartmouth Studies in Tropical Ecology 2014
CONSTRAINTS ON RESOURCE TRANSPORT IN ATTA COLOMBICA
MAYA H. DEGROOTE, ANN M. FAGAN, EMILY R. GOODWIN, ADAM N. SCHNEIDER, AND ZACHARY T.
WOOD
Faculty Editor: Matthew P. Ayres
Abstract: Organisms that transport and store resources can be constrained by their carrying ability. Atta colombica,
leaf-cutter ants, are colonial central place foragers that transport massive quantities of leaf fragments to their
mutualistic fungal partner (Agariceaea spp.) Unlike many organisms that transport resources, A. colombica alter the
physical form of their resource by cutting it into smaller pieces. We studied the factors that influence A. colombica
transport of leaf fragments. When harvesting leaves of variable mass per surface area, ants standardized leaf
fragment surface area relative to body mass. Therefore, average fragment mass varied by about two-fold. Atta
columbica may be constrained by leaf fragment surface area, which could be a function of handling ability. Ants
also carried loads that were only half as heavy as the apparent optimum, possibly because of physiological
constraints or selection acting on colonies rather than individuals.
Key words: Atta colombica, central-place foraging, leaf-cutter ants, load size determination
INTRODUCTION
Organisms that transport and store resources can
be constrained by their ability to carry those
resources. Transport strategies are varied and often
dependent on details of the specific resource.
Morphological structures used to transport
resources can vary from beaks to pincers, claws,
and pouches. Despite these derived morphologies,
organisms often have little to no control over the
physical form of the resource they carry. The size
or mass of the resource may limit transportation
efficiency and compromise other biological
imperatives such as mating and avoiding
predation.
Atta colombica, leaf-cutter ants, are somewhat
unusual in that they cut their own resource
fragments and therefore have control over the size
and mass of the resources they carry. As colonial
central place foragers, leaf-cutter ants transport
plant matter to a communal nest (Orians and
Pearson 1979). Leaf-cutter ants have a caste
system within which ants of different body sizes
perform different tasks. Mediae ants cut fragments
from the leaves of a variety of tropical plants and
carry them to the colony. These leaf fragments are
delivered to considerably smaller minim ants,
which cut the leaf fragments into smaller pieces
and feed them to Agaricaceae fungi. The fungi are
farmed as food for the leaf-cutter ants (Weber
1972). The rate of transportation of plant material
to the Agaricaceae fungi determines the amount of
fungus that can be farmed and influences the
growth and reproductive success of the colony.
We sought to understand the factors that
constrain A. colombica transport of leaf fragments.
Since leaf-cutter ants transport leaves of different
densities, leaf fragment surface area (henceforth:
leaf area) and leaf fragment mass (henceforth: leaf
mass) may vary. Both characteristics could affect
ease of leaf handling by ants. While heavier leaf
fragments may require more energy to carry, leaf
fragments with larger surface area may be difficult
to handle or have poor aerodynamics (Rudolph
and Loudon 1986). Standardization of a
characteristic (e.g., always carrying the ideal mass
of leaf) can optimize transport efficiency or
handling ability. If two characteristics are related
(e.g., fragment size and mass across different
densities), only one of these characteristics can be
standardized at once, and whichever characteristic
is the most constraining should be constant. We
tested whether ants standardize leaf area or mass
when harvesting from plants with variable leaf
densities. We also examined whether ants
maximize the rate of leaf mass transport.
METHODS
Data Collection
We conducted our study at the Sirena Biological
Station, Corcovado National Park, Costa Rica, on
4-5 February, 2014. We followed eight different
leaf-cutter ant trails to their source plants, where
we randomly collected twenty ants and their
77
Corcovado
respective leaf fragments. We selected single
worker ants carrying leaves, ignoring leaves
shared by multiple workers or those with minim
ants riding on top. We measured ant mass, leaf
mass, and leaf area. We calculated leaf density
using leaf mass and leaf area.
We also conducted speed trials on one trail of
leaf-cutter ants. We timed a single ant carrying a
leaf fragment one meter and calculated its
velocity. We then weighed each ant and its leaf
fragment.
Statistical Analysis
We employed a general linear model to evaluate
effects on leaf mass and leaf area of ant mass,
plant type, and their interaction. Visualizations of
leaf area and leaf mass across plant types used the
full model to adjust for effects of ant mass. We
also examined how adjusted leaf size and leaf
mass were related to mean leaf density (mass /
surface area).
From the speed trials, we calculated ant load
as mg of leaf carried per mg of ant body mass. We
fit a regression of ant speed to load, leaf mass, and
ant mass, then used the regression equation to
compare carrying efficiency (milligrams of leaf
carried per milligram of ant per second) to load
carried.
All analyses were performed using JMP® 11.0
statistical software (SAS Institute, Cary, NC).
RESULTS
Ant mass ranged by about 10-fold (≈ 3 to 30 mg)
and did not significantly vary among the antgroups harvesting different plant types (F7,147 =
1.95, P = 0.07). Larger ants carried leaves with
more surface area and greater mass (Equations 12, Figure 1B):
(1)
(2)
The mean area of leaf fragments ranged from
about 2 to 3 cm2, and the mean leaf density ranged
from about 6 to 10 mg/cm2. Across trails, leaf area
was unrelated to leaf density, but the mass of leaf
fragments increased with increasing leaf density
(Table 1, Figs. 2-3).
The load of leaf-carrying ants varied by about
8-fold (≈ 1 to 8 mg leaf/mg ant). The velocity of
leaf-carrying ants decreased over this range of
loads from about 4 cm/sec to about 2 cm/sec
(Equation 3, P = 0.0016, r2 = 0.34, Fig. 4):
(3)
The residuals of ant velocity vs. mg leaf/mg ant
(Fig. 4) showed no pattern with either leaf mass or
ant mass. Equation 4 is a rearrangement of
Equation 3 to give a measure of leaf transport
efficiency:
(4)
This equation has a maximum when Mleaf / Mant, or
load, is 7.53. This indicates that ants maximize the
mass of leaf per mass of ant that can be moved a
given distance per unit time when the fragment has
a mass 7.53 times that of the ant. Actual ants
carried much smaller loads than this (mean ± SD =
2.16 ± 1.31 mg/mg).
Table 1. Linear regression ANOVA results for the relationship between (1) ant mass
and leaf area and (2) ant mass and leaf mass across the eight plants that were
sampled. MS is the mean squares.
Surface area of leaf fragment
Source
df
Plant
7
1.72
2.85
0.009
Ant mass
1
19.87
32.73
< 0.0001
7
0.89
1.47
0.18
102
139
0.61
Plant*Ant mass
Error
78
MS
F
P
Mass of leaf fragment
MS
F
P
297
5.77
< 0.0001
1016
19.75
< 0.0001
1.98
0.06
51.4
Dartmouth Studies in Tropical Ecology 2014
A
7
Area of Leaf Fragment (cm2)
6
Plant 1
5
Plant 2
Plant 3
4
Plant 4
Plant 5
3
Plant 6
Plant 7
2
Plant 8
1
0
0
5
10
15
0
5
10
15
60
B
20
25
30
35
20
25
30
35
Mass of Leaf Fragment (mg)
50
40
30
20
10
0
Ant Mass (mg)
Figure 1. Larger ants carried leaf cuttings with (A) greater surface area and (B) greater mass. See text
for equations.
79
Corcovado
Area of leaf fragment (cm2 )
Figure 3. Mass of leaf fragments increased with
leaf density.
3.3
3.1
2.9
2.7
2.5
2.3
2.1
1.9
1.7
1.5
4
6
8
10
Mean leaf density (mg/cm2 )
Figure 2. Ants carried leaf fragments of the same
surface area regardless of leaf density.
Ant speed (m/s)
0.06
0.05
0.04
0.03
0.02
0.01
0
0
2
4
6
8
10
Load (mg/mg)
Figure 4. Ants carrying heavier loads traveled
more slowly. Ant load was defined as mg of leaf
mass carried per mg of ant body mass.
80
DISCUSSION
We tested whether leaf-cutter ants standardized
leaf mass or area when harvesting leaves of
variable mass per surface area. Leaf area remained
fairly constant across A. colombica loads (Fig. 2),
while leaf mass did not (Fig. 3). This indicates that
ants are more constrained by leaf area than leaf
mass. However, ants that carried heavier loads
traveled more slowly (Fig. 4). Furthermore, ants
carried much smaller leaf fragments than would be
expected if they were maximizing leaf transport
efficiency. A similar study on a different species
of leaf-cutter ants also found that they carried
pieces smaller than would be expected to
maximize efficiency (Frankel and Johnson 2012).
Conservation of leaf area by ants may be due
to several factors. For example, ant body size may
constrain leaf area if leaf cutting is a fixed action
pattern (Weber 1972). In fact, it seems that ants of
the same size tend to cut leaf fragments of the
same surface area because they position
themselves on the edge of a leaf and rotate around
a stationary point while cutting with their
mandibles. This does not preclude the possibility
that conservation of leaf area is optimal in terms of
handling and transport efficiency.
Analyses of ant velocity vs. loading indicated
that ants carried lighter than optimal loads. Ants
may be simply incapable of optimizing load.
Alternatively, their behavior may be optimal in a
way that our study did not capture. Lewis et al.
(2008) reported that ants adjust load sizes
depending on trail gradient, apparently to maintain
transportation efficiency under variable trail
conditions. It is possible that the ants in our study
were optimizing load for a different trail gradient
than where we recorded velocities. Another set of
possible explanations for the apparently
suboptimal loading of foraging ants follows from
leaf-cutter ants being eusocial insects. For Atta
columbica, fitness is an attribute of the colony, not
of the individual workers. It would not be
surprising if optimization for the colony has a
different solution than optimization for individual
workers (Burd 1996). For example, single ants
burdened by heavy loads can slow the rate of all
ants traveling in the same column, thereby
reducing overall colony efficiency. Thus,
apparently suboptimal behaviors on the individual
level may be due to physiological constraints or to
Dartmouth Studies in Tropical Ecology 2014
selection acting on colonies rather than
individuals.
Ayres for his assistance with the statistical
analyses.
ACKNOWLEDGEMENTS
We would like to thank the staff of Corcovado
National Park for their guidance as well as M.
AUTHOR CONTRIBUTIONS
All authors contributed equally.
LITERATURE CITED
Burd, M. 1996. Foraging performance by Atta
colombica, a leaf-cutting ant. American
Naturalist
148: 597-612.
Franco, N.B., N.S. Johnson. 2012. Load—velocity
tradeoffs and optimal foraging behavior in
Atta cephaloides, Dartmouth Studies in
Tropical Biology 124-131.
Lewis, O.T., M. Martin, T.J. Czaczkes. 2008.
Effects of trail gradient on leaf tissue transport
and load size selection in leaf-cutter ants.
Behavioral Ecology 19: 805-809.
Orians, G.H., and N.E. Pearson. 1979. On the
theory of central place foraging. D.J. Horn,
R.D.
Mitchell, G.R. Stairs (Eds.), Analysis of
Ecological Systems, Ohio State University
Press, Columbus pp 154-177.
Rudolph, S.G., and C. Loudon. 1986. Load size
selection by foraging leaf-cutter ants (Atta
cephalotes). Ecological Entomology 11: 401410.
Weber, Neal Albert. Gardening Ants: The Attines.
Vol. 92. Philadelphia: American Philos. Soc.,
1972.
81
Corcovado
THE SPATIAL DISPERSION OF IMMATURE AND ADULT TROPICAL CICADAS
ALEXA E. KRUPP, REBECCA C. NOVELLO, AND CLAIRE B. PENDERGRAST
Faculty Editor: Matthew Ayres
Abstract: Many organisms experience different selection pressures between different life stages, with consequences
for their physiology and behavior. Disparate selection pressures on juvenile and adult life forms may influence the
spatial distribution of individuals at different life stages. Like many insects, cicadas have very different ecology as
juveniles and adults. Nymphs feed underground on xylem fluids from tree roots and emerge as adults to feed, attract
mates, and reproduce in the forest canopy. We examined the spatial distribution patterns of nymphal and adult
cicadas in Corcavado National Park, Costa Rica. Mapping of nymphal exoskeleton abundance and adult calling
intensity in the forest revealed a strong tendency for aggregation in both life stages, but nymph abundance was not
related to adult abundance. Immatures and adults may rely on different resources even though both stages feed on
xylem fluids and show high spatial aggregation.
Key words: aggregation, cicadas, life stages, spatial distribution
INTRODUCTION
Selection pressures on organisms differ across life
stages, which may influence their spatial
distribution through time (Schluter et al. 1991,
Soler et al. 2013). In many insects, growth is the
measure of fitness relevant to immatures, while
adult life stages are focused on reproduction.
Saturinidae moths, for example, feed on leaves as
caterpillars, storing lipids that they will use
throughout their life. Adults, however, lack the
capacity to feed, instead devoting their energy
primarily to reproduction (van Nieukerken et al.
2011).
Cicadas also pass through distinct life stages
associated with varying selection pressures. After
females oviposit in trees, nymphs drop to the
ground and burrow into the soil, where they feed
on xylem fluids from tree roots. When they are
fully developed, flightless nymphs emerge from
the ground and molt to become adults capable of
flight. Adult males then aggregate in the canopy,
feeding on xylem fluids of above-ground branches
and chorusing to attract females for mating
(Young 1980). This basic life cycle is well
established, but the spatial distribution patterns of
cicadas and their dispersal abilities at different life
stages, particularly in the tropics, are not well
understood (Karban 1981, Williams and Simon
1995).
We examined the spatial distributions of
juvenile and adult male tropical cicadas,
considering possible drivers of these distributions.
If nymph development is driven by competition
82
for limited food resources, survival should be
higher in low-density areas, which would tend to a
uniform spatial distribution of nymphs. The effects
of competition could also cause nymphs in highdensity areas to be smaller than those in lowdensity areas. Alternatively, if nymph fitness is
driven by habitat quality, rather than limited by
food resources, nymphs should be most likely to
survive in high-quality habitats. In this case, we
would expect nymphs to be spatially aggregated
and larger nymphs to be found where densities are
high.
We also tested hypotheses of spatial
relationships between juvenile and adult cicada
populations. Since nymphs and adults both feed on
tree xylem, adults may have little incentive to
disperse from their site of emergence. Adult males
would then remain near the site of their nymphal
development to aggregate and chorus, perhaps
resulting in females oviposition near mating sites,
creating self-perpetuating assemblages around
high-quality resources (Oberdorster and Grant
2006). We could also see a spatial correlation
between sites of male chorusing and nymph
emergence if female dispersal is constrained.
Alternatively, if adults are mobile and their
resource needs are different, there might be no
correlation between juvenile and adult spatial
distribution.
METHODS
We conducted an observational study on 4-5
February 2014 at the Sirena Biological Station in
Dartmouth Studies in Tropical Ecology 2014
Corcovado National Park. We measured the
amplitude in decibels (dB) of cicada chorusing (a
putative measure of adult male cicada abundance)
every 200 paces along the Los Patos, Espaveles,
and Guanacaste trails between 08:00 and 15:00 on
5 February (N = 28 sites). We recorded elevation
and latitude/longitude of each site using Garmin
GPSmap 76CSx. We used the Real Time Analyzer
(RTA) iPhone application to estimate sound
energy at each site to the nearest half decibel. We
repeated decibel measurements on the Espaveles
trail three times (N = 11 sites) – once on 4
February (14:30 - 15:30) and twice on 5 February
(9:30 - 10:30 and 12:30 - 13:00). We also used the
RTA application to estimate the peak frequency of
cicada calls.
We divided each sample site into three
sections, and each investigator searched one
section for cicada exoskeletons for three minutes.
We took haphazard subsamples of exoskeleton
length at sites with fewer than 10 exoskeletons
(low-density) and more than 40 exoskeletons
(high-density). We measured 5-8 exoskeletons at
each of four low-density plots (total of 18
exoskeletons) and 8-10 exoskeletons at each of
three high-density plots (total of 19 exoskeletons).
Statistical Analyses
We converted decibels to sound intensity in
milliwatts/m2.
(1) dB = 10 ´ log10 (
measurement times). We compared the variation
among sites in the abundance of nymphs and adult
males in terms of the relative difference between
the 10th and 90th percentiles of exoskeleton
abundance and sound intensity. We also evaluated
the dispersion of exoskeletons using the ratio of
variance to mean for the sample plots
(variation/mean = 1 for random dispersion, < 1 for
uniform dispersion, > 1 for clumped dispersion).
We used regression analyses to test for a
relationship between sound intensity and
exoskeleton frequency. We used a two-tailed t-test
to compare exoskeleton sizes between randomly
selected low-density and high-density sites.
Finally, we used a Monte Carlo randomization test
to evaluate whether sound intensity or
exoskeletons were more similar between adjacent
sites (separated by 200 paces) than among sites in
general. For the randomization tests, we created a
frequency distribution of the sum of squared
differences between each pair of adjacent sites
over 1000 randomizations with the vector of
observed values randomly distributed among sites.
We compared the actual sum of squared
differences among adjacent sites to this spatially
randomized distribution. We conducted all
analyses in JMP 10 (SAS Institute, Cary, NC),
Microsoft Excel 2008, and R.
I1
), I0 = 10 -9mW/ m2
I0
The intensity of a sound wave represents the
amount of energy transported through an area per
unit time. This range of sound intensities
detectable to the human ear is so large that a
logarithmic scale is often used to represent sound
intensities. The lower threshold of human hearing
(10-9 mW/m2) is assigned an intensity of 0
decibels. A sound 10x times more intense than
another sound is 10x the decibels (The Physics
Classroom 2014). We chose to represent cicada
calling in mW/m2 rather than decibels to
accurately represent the differences in energy
expended by adult male cicada communities rather
than the sound landscape detected by the human
ear.
We tested the repeatability of sound intensity
across sites with a random effects ANOVA of
measurements on the Espavales Trail (11 sites x 3
Figure 1. Spatial patterns in sound intensity (chiefly
from cicadas) were repeatable through multiple
observation periods.
83
Corcovado
between 10th and 90th percentiles. Based on a
variation-to-mean ratio (VMR) of 25, exoskeleton
spatial distribution was aggregated at the site level
(Cox and Lewis 1966). Neither sound intensity nor
exoskeletons were more similar between adjacent
sites than expected by chance (Fig. 3). Local
exoskeleton abundance did not predict sound
intensity from cicadas (P = 0.18: Fig. 4, 5).
Figure 2. Exoskeleton size (± SE) was almost
identical between sites of low and high local
abundance.
RESULTS
Points in space had repeatable patterns in noise.
60% of variation in sound intensity was
attributable to location (Fig.1, random effects
ANOVA). Peak frequency recorded at each site
was 3.15 kHz. Hundreds of adult cicadas were
observed feeding in Cedrela odorata trees at
multiple sites between approximately 35 and 65 m
from the ground. Average exoskeleton length was
24.78 ± 2.45 mm and did not vary with
exoskeleton density (t = 0.16, P = 0.87, df = 33.26;
Fig. 2).
Sound intensity increased approximately 70fold between 10th and 90th percentiles, and
abundance of exoskeletons ranged from 0 to 48.8
DISCUSSION
Male cicada calling displayed consistent trends of
sound intensity over time; while average sound
intensity varied somewhat between observation
periods, the locations of relative high and low
intensity remained consistent (Fig. 1). This
suggests that groups of male cicadas do not
modify their position or chorusing intensity
relative to others in response to short-term
environmental variations. These findings justify
our choice of sound intensity as a metric for adult
male cicada abundance. As cicadas of both sexes
are attracted to congregational male chorusing
(Karban 1981), it is also likely that cicada song
indicates the presence of both male and female
adults.
We frequently found adult cicadas, high sound
intensity, and high exoskeleton abundance
concurrently in areas containing Cedrelo
odorata, a species of canopy tree. C. odorata is
Figure 3. Results of randomization tests for adjacent sites being more similar than by chance in sound
intensity (left) and exoskeleton. In neither case was there evidence of adjacent sites being similar. Arrows
represent actual values relative to random distribution.
84
Dartmouth Studies in Tropical Ecology 2014
Figure 4. Exoskeleton abundance and sound
intensity were not correlated.
found in well-drained soils with moderate to high
light availability, and therefore may indicate soil
and canopy conditions conducive to both adult and
nymph growth and survival (ISSG 2006).
Exoskeleton length did not vary between highand low-density areas (Fig. 2). We observed little
variation in exoskeleton size and a consistent peak
frequency of cicada song across all sites,
suggesting that our study system primarily
contained a single species (Sueur 2002, Janzen
1983). If exoskeleton size indicates nymph habitat
quality, stable exoskeleton length across plots with
variable local densities suggests that nymphs are
not resource-limited at any density. However,
nymph abundance may be a more appropriate
indicator of habitat quality than individual size.
Exoskeleton abundance and sound intensity
varied greatly through space, suggesting
aggregations in both nymphs and adults (Fig. 5).
Adult aggregation has been established as an
essential component of mating behavior while
nymph aggregation may be explained by patchy
resource availability for nymphs.
The locations of high abundance for adults and
nymphs were not obviously correlated (Fig. 5).
This was surprising because cicadas are weak
fliers and energy requirements and predation
associated with dispersal may be high (Karban
1981). Additionally, habitat quality for nymphs
and adults may be driven by similar factors due to
the shared xylem resource (Yang 2006). The lack
of correlation may indicate that adult males are reaggregating in the canopy following emergence.
Favorable canopy areas for adult male cicadas
may be located in different areas than high quality
nymphal environments. Cicada chorusing behavior
has been linked to environmental conditions such
Figure 5. Spatial patterns in cicada sound intensity (light gray) and exoskeleton abundance (dark gray).
Circe size represents low to high values.
85
Corcovado
as light availability (Yang 2006), while the drivers
of oviposition site selection by female cicadas are
largely unknown. Further investigation of the
biotic and abiotic factors affecting adult cicada
distribution may shed light on the relationship
between nymphal and adult populations.
Nymph and adult male cicadas both display
highly aggregated spatial dispersion in the forest at
the site level. As neither exoskeletons nor sound
intensity were aggregated across sites (Fig. 3),
aggregation appears to occur only at a finer scale.
Location and spacing of aggregations may reflect
habitat quality, mating pressure, or an interaction
of factors at these sites. However, ultimate drivers
of cicada distribution cannot be fully understood
without further analysis of the scale at which these
aggregations occur and the environmental features
associated with them. A deeper understanding of
what factors influence organisms’ spatial
distributions over the course their lifetimes could
provide better predictions for how a changing
environment will affect organisms’ distributions.
LITERATURE CITED
Aanen, D. K., H. H. D. Licht, A. J. M. Debets, N.
A. G. Kerstes, R. F. Hoekstra, and J. J.
Boomsma. 2009. High symbiont relatedness
stabilizes mutualistic cooperation in fungusgrowing termites. Science 326:1103-1106.
Calsbeek, R., L. Bonvini, and R. M. Cox. 2010.
Geographic variation, frequency-dependent
selection, and the maintenance of a femalelimited polymorphism. Evolution 64:116-125.
Chazdon, R. L., R. W. Pearcy, D. W. Lee and N.
Fetcher. 1996. Photosynthetic responses of
tropical plants to contrasting light
environments. Pages 5-55 in Mulkey, S., R.
L. Chazdon and A. P. Smith, editors. Tropical
forest plant ecophysiology. Chapman and
Hall, New York.
Kozlowski, T. T., P. J. Kramer, and S. G. Pallardy.
1991. The physiological ecology of woody
plants. Academic Press, New York.
Darwin, C., and A.R. Wallace. 2012. An
hypothesis to explain the diversity of
understory birds across Costa Rican
landscapes. Dartmouth Studies in Tropical
Ecology 2012, in press.
Wilkinson, B. L., J. H. M. Chan, M. N.
Dashevsky, D. L. Susman, and S. E. Wengert.
2009. Maternal foraging behavior and diet
composition of squirrel monkeys (Saimiri
oerstedii). Dartmouth Studies in Topical
Ecology 2009, pp. 141-144.
86
ACKNOWLEDGEMENTS
We would like to thank Matt for his wisdom and
guidance in the forest and out and Hannah ter
Hofstede for sharing her extensive knowledge of
tropical insects and sound measurements.
AUTHOR CONTRIBUTIONS
All authors contributed equally. Becca ran analysis
in R.
Dartmouth Studies in Tropical Ecology 2014
SPATIAL DISTRIBUTION OF LYCOSIDAE SPIDERS WITHIN AND ACROSS HABITATS
ALEXA E. KRUPP, AARON J. MONDSHINE, CLAIRE B. PENDERGRAST,
AMELIA L. RITGER, AND ADAM N. SCHNEIDER
Faculty Editor: Matthew P. Ayres
Abstract: Animals should preferentially select habitats where their fitness is maximized. Good habitats may offer
high quality food resources, favorable abiotic conditions, or reduced vulnerability to predation. Ecological theories
such as ideal free distribution, ideal despotic distribution and optimal foraging theory can predict common patterns
of animal distribution. We considered the relative capacity of these three theoretical models to explain Lycosidae
spider spatial distribution patterns at the micro and macrohabitat scale at the La Selva Biological Station, Costa
Rica. We observed smaller spiders at higher densities in forest areas, and bigger spiders at lower densities in open
fields. We found no relationship between body size and spider density at any spatial scale. Our findings suggest that
optimal foraging determines community structure within a macrohabitat. Prey availability within a plot influenced
spider density, providing evidence for ideal free distribution theory on the microhabitat level.
Key words: Ideal despotic distribution, ideal free distribution, Lycosidae, macrohabitat, microhabitat, optimal
foraging theory
INTRODUCTION
Organisms achieve maximal fitness by
preferentially selecting habitats best suited to their
morphology, life history, and resource
requirements. Habitat selection should maximize
opportunity for resource acquisition while
minimizing the risks of predation. As
environmental conditions vary with location, time,
and scale, individuals seek out habitats best suited
to their physiology and needs.
Various ecological models seek to identify and
predict distribution patterns observed throughout
the natural world. Under the ideal free distribution
individuals distribute themselves among habitats
in proportion to the quality of each habitat
(Sinervo 2006). The ideal free distribution theory
assumes that habitat quality decreases with
increasing population density due to reduced per
capita resource availability, and that all individuals
are competitively equal (Sinervo 2006).
Alternatively, under the ideal despotic
distribution, the best competitors dominate and
control access to highest quality habitats, resulting
in the exclusion of less competitive individuals
from areas of high resource availability (Sinervo
2006). Finally, optimal foraging theory posits that
different foraging strategies allow individuals to
maximize energy acquisition while minimizing
risk of predation in response to environmental
conditions (Pyke et al. 1977). Optimal balance of
foraging strategy and predator avoidance often
varies with organism size or life stage. For
example, many reef fishes occupy protected
mangroves as juveniles before moving onto the
open reef as adults with higher energy demands
and reduced vulnerability to predation
(Nagelkerken et al. 2000).
We compared the population densities and
body sizes of wolf spiders (family Lycosidae)
within and across two habitat types, which allowed
us to assess the spider population’s spatial
distribution on both the macro and microhabitat
level. Lycosidae are generalist predators that
consume prey of variable size and type, from ants
to small vertebrates (Marshall 1996; McCormick
and Polis 2008). Body size varies widely among
individuals and species, resulting in variable
energetic requirements and prey capture efficiency
(Birkhofer et al. 2006).
If Lycosidae distribute themselves
accordaning to ideal free distribution, spider
density would increase with microhabitat quality,
while mean body size would remain constant
across microhabitats. If instead, spiders are
distributed accordaning to the ideal despotic
distribution, higher quality microhabitats would be
occupied by larger spiders capable of excluding
smaller spiders (Birkhofer et al. 2006), while
spider density would remain constant. Finally, if
spider populations displayed dynamic optimization
between size classes, mean body size would vary
between macrohabitats.
87
La Selva
METHODS
Data Collection
We assessed Lycosidae density and habitat
suitability in La Selva National Park, Costa Rica
between 17:00 and 21:00 on 15-16 February 2014.
We selected three sites in open fields and three
sites in secondary forest. For each site, we
randomly sampled eight 1 x 3 meter plots along a
60-meter transect and measured spider abundance
and spider body length from head to thorax. We
found spiders by scanning a plot using a headlamp
to locate eye-shine, an established method for
locating Lycosidae spiders (Kaston 1978). We also
spent 60 seconds counting the number of insects,
which are potential prey items, inside each plot.
Figure 1. Spiders were smaller on average in
forest sites than field sites but field sites also
varied in body size.
Statistical Analysis
We performed an Analysis of Variance (ANOVA)
for spider body size and density with respect to
habitat type. We regressed spider density with
respect to mean body size within each habitat type.
We also performed an ANOVA to compare prey
density across habitat types and for average body
size with respect to prey density and habitat.
Finally, we evaluated spider density with respect
to prey density, habitat type and their interaction
using a generalized linear model. All analyses
were performed using JMP® 11.0 statistical
software (SAS Institute, Cary, NC).
RESULTS
We found Lycosidae spiders with head to thorax
lengths ranging from 1.4 to 21.0 mm from a
number of different genera (N = 168), including
Figure 2. Spider density was higher in the forest
than in the field.
Figure 3. Spider body size did not predict density in either habitat type.
88
Dartmouth Studies in Tropical Ecology 2014
Figure 4. Spider density was positively correlated
with prey density and was higher in the field than the
forest.
Sosippus sp. (adults vary between 2.5 and 18.5
mm), Pardosa sp. (adults vary between 4.5 and 9.5
mm), and Lycoa sp. (adults range from 9.0 to 35.0
mm; Kaston 1978).
Spider size range was consistent across habitat
type, but spiders were smaller on average in the
forest than in the field (F5, 157 = 4.51, P = 0.004).
Mean body size also varied significantly within
each habitat (F5, 157 = 4.51, P = 0.002: Fig. 1).
Spider density was significantly higher in the
forest than in the field (F5, 42 = 4. 22, P = 0.001:
Fig. 2). Although there was a higher density of
smaller spiders in the field and a lower density of
larger spiders in the forest, there was no
relationship between body size and spider density
within each habitat (slope = -0.09 ± 1.5, P = 0.59,
r2 = 0.01 and slope = 0.09 ± 2.2, P = 0.67, r2 = 0.01
respectively: Fig. 3). Therefore, body size did not
predict spider density within or across habitats.
Prey density was over three times higher in the
field than in the in forest (F5, 38 = 2.97, P = 0.003).
Spider density was positively correlated with prey
density (F3, 40 = 4.53, P = 0.03) and was higher in
the field than the forest (F3, 40 = 4.53, P = 0.001:
Fig. 4). The slope of spider density versus prey
density did not differ between habitats (F3,38 =
1.84, P = 0.23).
DISCUSSION
We tested whether the spatial distribution of
Lycosidae spiders at the micro and macrohabitat
scale was a better match with the ideal free
distribution, ideal despotic distribution theory, or
optimal foraging theory.
Within both forest and field sites, we observed
greater spider density in plots with more prey
items (Fig. 4), indicating that resource availability
influences spider distribution on the microhabitat
scale. No relationship between spider body size
and prey density was observed, suggesting that
competitive exclusion by size was not occurring in
areas of high resource availability. These results
jointly support the ideal free distribution theory as
an explanation for spider distribution on a
microhabitat scale.
On a macrohabitat scale, patterns of spider
size and density did not fit the assumptions of
ideal free distribution observed on the
microhabitat level. Larger, less densely distributed
spiders were found in the field sites, while forests
contained greater densities of smaller spiders (Fig.
1, Fig. 2), despite higher prey availability in the
field sites. This pattern indicates that optimal
foraging theory may drive spider spatial
distribution across habitat types, as spiders in
different size classes likely perceive the value of
the two habitats differently and therefore do not
compete over the same areas.
Prey abundance was not the sole driver of
spider distribution. Risk of predation, among other
factors, may outweigh potential fitness gains
offered by higher prey abundance in an
environment. Although field sites had more prey
available than did forest sites, capture success may
be lower or the cost of increased vulnerability to
predation may be higher for smaller spiders in the
open environment. Thus, environmental conditions
in the forest might be more optimal for smaller
spiders. Our methodology did not fully capture
variations between forest environments with
regards to predator and prey presence, as well as
with substrate types, and it is likely that these and
other differing macrohabitat conditions also
strongly affect spider fitness.
Organisms distribute themselves in response
to environmental conditions according to different
patterns depending on scale. Recognizing how
these patterns interact with one another is
important to understanding interactions between
populations and their environment.
89
La Selva
ACKNOWLEDGEMENTS
We thank Matthew Ayres, Jessica Trout-Haney,
and Vivek Venkataraman for their assistance.
AUTHOR CONTRIBUTIONS
All authors contributed equally in the data
collection and writing process.
LITERATURE CITED
Birkhofer, K., Henschel, J.R., and S. Scheu. 2006.
Spatial-pattern analysis in a territorial spider:
evidence for multiscale effects. Ecography 29:
641-648.
Kaston, B.J. 1978. How to know the spiders.
Brown Company Publishers, Dubuque, IA.
Kacelnik, A., Krebs, J.R., and C. Bernstein. 1992.
The ideal free distribution and predator-prey
populations. Tree 7: 50-55.
Marshall, S.D. 1997. The ecological determinants
of space use by a burrowing wolf spider in
xeric shrubland ecosystem. Journal of Arid
Environments 37: 379-393.
McCormick, S., and G.A. Polis. 2008. Arthropods
that prey on vertebrates. Biological Reviews
57: 29-58.
Nagelkerken, I., van der Velde, G., Gorissen,
M.W., Meijer, G.J., Van Hof, T., and C. den
Hartog. 2000. Importance of Mangroves,
Seagrass Beds and the Shallow Coral Reef as
a nursery for Important Coral Reef Fishes,
Using a Visual Census Technique. Estuarine,
Coastal and Shelf Science. 51: 31-44.
Pyke, G.H., H. R. Pulliam, and E.L. Charnov.
1977. Optimal foraging: a selective review of
theory and tests. Quarterly Review of Biology
52: 137-154.
Samu, F., Sziranyi, A., and B. Kiss. 2002.
Foraging in agricultural fields: local ‘sit-andmove’ strategy up to risk averse habitat use in
a wolf spider. Animal Behavior 66: 939-947.
Sinervo, B. 2006. The Ecology of Foraging.
University of California, Santa Cruz. Santa
Cruz, CA.
90
Dartmouth Studies in Tropical Ecology 2014
THE RICH GET REDDER: DELAYED GREENING IN WELFIA REGIA PALMS
ALLEGRA M. CONDIOTTE, ANN M. FAGAN, NATHANAEL J. FRIDAY, ELLEN R. PLANE, AND
ZACHARY T. WOOD
Faculty Editor: Matthew P. Ayres
Abstract: Plants must allocate resources efficiently to maximize growth, defense, and reproduction. Herbivory has
selected for a variety of antiherbivore defenses. One of these defenses is delayed greening in new leaves, which
reduces attractiveness to herbivores until a leaf has matured and become less palatable. However, leaves that delay
greening have reduced photosynthetic capacity, making this form of defense energetically costly. We examined
whether the duration of delayed greening in Welfia regia palms is fixed or is plastic with respect to previous
hervivory, light availability, or plant size. The duration of leaf redness increased with plant size, but was not related
to light availability or herbivory. Evidently W. regia does not adjust the developmental program for young leaves
based on previous herbivory or light availability. An increase in metabolic capacity with size appears to be the major
factor driving duration of delayed greening. Probably large palms can afford the cost of reduced photosynthetic
capacity in their youngest leaf in order to avoid herbivory, while small palms require more photosynthetic area
earlier in order to develop. Our results suggest that large individuals of W. regia have enough resources to
simultaneously grow, reproduce, and defend their young leaves against herbivory.
Keywords: delayed greening, herbivore defense, resource allocation, Welfia regia
INTRODUCTION
Plants must manage limited resources for growth,
reproduction, and herbivory defense, all of which
are essential to fitness. Plants should allocate the
highest proportion of resources to whatever
particular attribute will maximize long-term
fitness. In environments with high rates of
herbivory, plants may forgo some amount of
growth and reproduction to protect tissue from
herbivory. As energy and nutrients are limiting,
plants must maintain a balance between using
energy to support expensive defenses and growing
energetically but losing productivity to herbivores.
Rates of herbivory are higher in tropical
environments than in the temperate zone, and most
damage to plant tissue is caused by herbivory
(Coley and Barone 1996). Young leaves, which
are less tough than mature leaves during the rapid
expansion of early growth, are particularly
vulnerable to herbivory (Dominy et al. 2002).
Plants in the tropics have evolved a variety of
antiherbivore defenses to protect these younger
leaves. For example, delayed greening postpones
the input of chlorophyll, nitrogen, energy, and
rubisco into young leaves, keeping leaves red and
keeping them low in nutritive value to potential
herbivores until leaves become tougher and less
palatable at maturity (Kursar and Coley 1992).
Delayed greening reduces photosynthetic
capacity (Kursar and Coley 1992). Leaves with
delayed development of chloroplasts absorb only
18-25% of maximum possible light, as opposed to
normally developed leaves, which absorb 80%
(Kursar and Coley 1996). This loss of
photosynthetic ability is costly, and may limit the
number of new leaves that exhibit delayed
greening.
We studied the factors affecting delayed
greening in the tropical palm Welfia regia
(Arecaceae). We tested whether delayed greening
in W. regia is inducible (activated in response to
an external stimulus) by examining whether plants
experiencing increased herbivory had more
delayed greening. We also tested whether light
availability or metabolic capacity (by size) limited
delayed greening. If photosynthetic needs limit
delayed greening, W. regia with high amounts of
epiphyte cover on leaves and low canopy openness
should be less likely to delay leaf greening
because they have higher need to photosynthesize.
If plant metabolic capacity (e.g., total pools of
carbohydrates and nutrients) is limiting, small
palms should exhibit less delayed greening.
METHODS
We sampled 36 W. regia palms along the Camino
Experimental Sur and Sendero Tres Rios trails at
La Selva Biological Station, Costa Rica on 15-16
February, 2014. We measured plant height at the
tallest point using a hypsometer and measured
91
La Selva
width at the longest horizontal axis and the
perpendicular horizontal axis using a measuring
tape.
On each palm, we ranked the five newest
leaves from youngest (A) to oldest (E) based on
their proximity to the center of the plant. We then
randomly selected one leaflet on each of these
leaves and inspected it for herbivore damage and
epiphyte cover. Two observers independently
ranked herbivore damage on a scale from 0 to 5 (0
= no damage, 1 = small hole, 2 = small holes, 3 =
large hole, 4 = large holes, 5 = most of leaf
damaged). We also ranked epiphyte cover on a 0-5
scale (0 = no epiphytes, 1 ≈ 10% epiphyte cover, 2
≈ 25% epiphyte cover, 3 ≈ 50% epiphyte cover, 4
≈ 75% epiphyte cover, 5 ≈ 100% epiphyte cover).
We evaluated the correlation of scores between
observers and then averaged the rankings of both
observers for subsequent analyses of herbivore
damage and epiphyte cover.
We dried each of the randomly selected
leaflets from the three youngest leaves for 24
hours in a drying oven. We subtracted the final dry
mass from the initial wet mass to determine
percent water content of each leaf.
We took four canopy photographs above each
plant using a Canon PowerShot SX130IS and
analyzed percent canopy cover in ImageJ 764
software. We also took flash photographs of a
random leaflet on the plant’s youngest leaf and
analyzed its redness using Photoshop 5.0 (Adobe
Systems Incorporated, San Jose, CA). We
compressed a cropped section of each leaf
photograph into one pixel, recorded its red, blue,
and green saturation, and used Equation 1 to
determine leaf redness (Barber et al. 2000;
Freschknecht 1993).
R
(R + G + B)
3
Equation 1
Statistical Analyses
We used principal component analyses (PCA) to
derive a single measure for each plant of size,
herbivory, and epiphyte coverage. For our size
measure, we took the first principal component of
a PCA of plant height, maximum width, and
orthogonal width (see loadings in Table 1; 90% of
variation explained). For herbivory, we took the
first axis of a PCA on herbivory scores for the five
92
leaf ages (loadings for leaves A-E = -0.49, -0.48, 0.49, -0.29, and -0.45, respectively; 33% of
variance explained). We followed the same
procedure for epiphyte coverage (loadings for
leaves A-E = -0.33, -0.45, -0.52, -0.46, and -0.46,
respectively; 40% of variance explained).
Henceforth, the terms “size,” “herbivory,” and
“epiphyte coverage” all refer to the first axis of the
corresponding PCA. We performed analysis of
variance (ANOVA) to compare water content of
leaves of different ages. We conducted a
regression comparing leaf water content with leaf
redness of the youngest leaf. To assess effects
other than age, we used the residuals from this
regression for further analysis.
We fit regressions of residual redness to size,
herbivore damage, epiphyte cover, and canopy
openness and used ANOVAs to test whether each
regression was significant.
RESULTS
Water content varied as a function of leaf age
(F2,108 = 34.9, P < 0.0001). The youngest leaf
contained more water than the second-youngest
and third-youngest leaves (Fig. 1). Thus, water
Table 1. PCA results for palm size
Mean ± S.D.
PC-1
loadings
Maximum
Height
3.3 ± 1.4
0.55
Maximum
Width
4.5 ± 2.7
0.59
Orthogonal
Width
3.6 ± 2.3
0.59
Metric
content served as a proxy for leaf age. leaf redness
increased with water content (Fig. 2), indicating
that W. regia leaves become less red with age
(slope = 0.029 ± 0.019, F1,33 = 19.7 P < 0.001, r2 =
0.37).
Canopy cover, herbivory, and epiphyte cover
were all unrelated to residual leaf redness.
However, residual redness increased with palm
size, indicating that larger palms retain the redness
in their leaves for longer into leaf maturation than
smaller palms (slope= -0.050 ± 0.175, F1,33=8.1,
P= 0.008, r2= 0.20; Fig. 3).
Dartmouth Studies in Tropical Ecology 2014
Figure 3. The young leaves of larger plants were
redder than the young leaves of smaller plants.
Figure 2. Redness (Equation 1) increased with water
content, indicating that younger leaves were redder
than older leaves.
The youngest leaves (leaf A) had the lowest
herbivory rankings and usually had less than 10%
herbivory. Older leaves tended to have more
surface area loss due to herbivory, often with
multiple large holes (Fig. 4).
DISCUSSION
Delayed greening reduces herbivory on young
leaves. As new leaves age, they transition from red
to green, although not necessarily at the same rate.
We found that greening was delayed longer in
larger plants but was not related to light
availability, previous herbivory, or epiphyte cover.
The greater metabolic capacity resource pools
in larger W. regia could explain why new leaves
on larger plants stayed redder longer than those on
smaller plants (Figure 3). Larger plants have
greater photosynthetic surface area, more access to
light in the understory, and greater energy and
nutrient reserves. Therefore, they can invest more
energy in creation of a red leaf and afford the cost
of reduced photosynthetic efficiency in their
youngest leaf for longer.
This pattern suggests a shift in W. regia
resource allocation priorities as plants develop.
Plants at earlier life stages prioritize quick growth
and establishment despite increased risk of
herbivory, while older, established palms
minimize herbivory at the cost of diminished
growth rate. Mortality in young tropical palms that
do not engage in delayed greening is significantly
higher than in older palms and in young palms that
delay greening (Numata et al. 2004), indicating
that young plants, which have limited resources to
allocate to defense, are more vulnerable in general
than their older counterparts. Young plants face
high mortality rates due to their location in the
understory. There may be an age structure
bottleneck, in which mortality is high in young
plants due to herbivory, low light availability, and
disease, from which the few surviving plants are
released upon reaching a certain age.
There was no relationship between duration of
redness and herbivory, epiphyte cover or canopy
cover. Because delayed greening did not change
based on any of these factors, it appears that this
defense is not a plastic response to environmental
conditions, but is instead a fixed ontogenetic
trajectory. Keeping the youngest leaf red as long
as possible may be necessary for survival in an
environment with high rates of herbivory.
Plants face tradeoffs in resource allocation for
growth, defense and reproduction. Plants in high
resource systems have their highest realized
growth with minimal defense allocations because
they can afford to sacrifice some resources (Coley
et al. 1985). Young plants that devote all their
resources to defense will not grow big enough to
establish themselves. However, new growth fails
to benefit a plant if it is all consumed by
herbivores. Our finding that younger W. regia
93
La Selva
Figure 4. The youngest leaves had little to no herbivory, while older leaves had higher levels of herbivory. Leaf
A is the youngest, and E the oldest (leaf ages determined by position within the whorl).
invest fewer resources in delayed greening
suggests that growth to a certain size threshold is
the limiting factor for long-term plant success,
after which the plant has enough resources to
continue growth and reproduction while
simultaneously investing in greater herbivore
defense.
ACKNOWLEDGEMENTS
We thank Matthew Ayres, Jessica Trout-Haney
and Vivek Venkataraman for their guidance and
LITERATURE CITED
Barber, I., Arnott, S.A., Braithwaite, V.A.,
Andrew, J., Mullen, W., and Huntingford,
F.A. 2000. Carotenoid based sexual coloration
and body condition in nesting male
sticklebacks. Journal of Fish Biology 57: 777790.
Coley, P.D. and Barone, J.A. 1996. Herbivory and
plant defenses in tropical forests. Annual
Review of Ecology and Systematics 27: 305335.
Coley, P.D., Bryant, J., and Stuart Chapin III, F.
1985. Resource availability and plant
antiherbivore defense. Science 230: 895-899.
Dominy, N.J., Lucas, P.W., Ramsden, L.W., RibaHernandez, R., Stoner, K.E., and Turner, I.M.
2002. Why are young leaves red? Oikos
98:163-176.
94
direction in the experimental design and writing
process.
AUTHOR CONTRIBUTIONS
All authors contributed equally. A. Fagan, N.
Friday, E. Plane, and Z. Wood collected field data
and A. Condiotte conducted water content
analysis.
Frischknecht, M. 1993. The breeding colouration
of male three-spined sticklebacks
(Gasterosteus aculeatus) as an indicator of
energy investment in vigour. Evolutionary
Ecology 7:439- 450.
Kursar, T.A., and Coley, P.D. 1992. Delayed
greening in tropical leaves: an antiherbivore
defense? Biotropica 24:256-262.
Kursar, T.A. and Coley, P.D. 2006. The
consequences of delayed greening during leaf
development for light absorption and light use
efficiency. Plant, Cell & Environment 15:901909.
Numata, S., Kachi, N., Okuda, T., and Manokaran,
N. 2004. Delayed greening, leaf expansion,
and damage to sympatric Shorea species in a
lowland rain forest. Journal of Plant Research
117:19-25.
Dartmouth Studies in Tropical Ecology 2012
LOG PREFERENCE IN LEPIDOPHYMA FLAVIMACULATUM: SHELTER, HUNTING GROUND,
OR BOTH?
MAYA H. DEGROOTE, LARS-OLAF HÖGER, AND REBECCA C. NOVELLO
Faculty Editor: Matthew P. Ayres
Abstract: Many animals use structures found in their environments to provide protection from predators, and these
structures are sometimes also used as hunting grounds. The secretive yellow-spotted tropical night lizard,
Lepidophyma flavimaculatum, is often found under logs, and little is known about its activity. We tested whether
tropical night lizards use logs as hunting grounds or solely as shelters. We compared various log and soil attributes
that might influence arthropod abundance between occupied and unoccupied logs and tested for relationships
between these attributes and lizard condition. We found that tropical night lizards use logs primarily as hiding
places. In addition, tropical night lizard condition decreased with log and soil temperature, indicating that tropical
night lizards benefit from having a lower metabolic rate as they rest beneath logs during the day. This suggests that
tropical night lizards are nocturnal, foraging at night and using the logs as shelters during the day. Understanding the
conditions that influence habitat suitability provides insight into how animals position themselves in their
environment.
Keywords: Lepidophyma flavimaculatum, log, shelter
INTRODUCTION
Throughout their lives, animals must make
decisions about where to spend their time. Many
animals use structures in their environments as
shelters to reduce risk of predation. Some birds,
for example, excavate or co-opt previously
excavated chambers in trees for nesting. These
cavity-nesting birds are protected from both
environmental stressors and predation as they rest
at night. Sometimes, an organism’s shelter also
acts as a hunting ground. A combined shelter and
hunting resource is beneficial as it reduces an
organism’s need to travel and risk exposure to
predation.
Lizards escape predation using a wide array of
mechanisms, including crypsis, speed, taildropping, and hiding under environmental
shelters. The parthenogenetic yellow-spotted
tropical night lizard, Lepidophyma
flavimaculatum, tends to hide beneath logs and
within rock crevices (Cooper, 1994). As such,
herpetologists are unsure if tropical night lizards
are nocturnal or diurnal (Savage 1992), raising
questions about their foraging strategies and daily
activities. Much of what we know about the
tropical night lizard is inferred from the ecology of
closely related species, many of which are ambush
predators (Cooper 1994). Ambush predators hide
in plain site or conceal themselves using
environmental structures where they wait to strike
when prey approach. If night lizards are in fact
ambush predators, a log that can conceal them
may serve as both a shelter and a hunting ground,
providing them access to arthropod prey.
We investigated whether tropical night lizards
use logs solely as shelter or as both shelter and
hunting grounds. If lizards use logs as both shelter
and hunting grounds, factors improving their
hunting success should be greater beneath
occupied than unoccupied logs. We used several
soil and log attributes that might increase
arthropod abundance, including soil moisture,
organic horizon depth, and soil and log
temperature, as proxies for hunting success. Under
this hypothesis, attributes promoting arthropod
abundance should also be positively correlated
with lizard condition. However, if lizards use logs
solely as shelters, there should be no difference in
these attributes between occupied and unoccupied
logs. Under this hypothesis, there should likewise
be no relationship between these attributes and
lizard condition.
METHODS
We surveyed logs in a 3000 m2 area of the
arboretum and within 10 m of the sides of 500 m
of the Sendero Tres Rios trail at the La Selva
Biological station on 15-16 February, 2014
between 10:00 and 16:00. We overturned logs that
appeared to be structurally viable for lizard
occupation, which we defined as having a space of
at least an inch between the log and the soil. We
95
La Selva
noted the presence or absence of lizards under
each log. We surveyed a total of 15 logs without
lizards (7 in the arboretum, 8 along the trail) and 8
logs with lizards (4 in the arboretum, 4 along the
trail). We measured the snout-to-vent length
(SVL) and body mass of all lizards found. We also
noted the presence or absence of arthropods under
each log and took measurements of several log
and soil conditions that might affect arthropod
success and abundance (i.e., soil moisture,
temperature of both the soil and underside of the
log, log decomposition, area under the log, log
volume, and soil organic horizon depth). We
placed soil samples from under each log in a
drying oven at 70˚C for 20-25 hours and
calculated soil moisture as the percent moisture by
mass. Temperature was determined using a Raytek
Raynger MX temperature gun. Log decomposition
was estimated on a scale of 1-5 with 5 being
highly decomposed.
Statistical Analyses and Modeling
We used two-tailed t-tests to compare log and soil
attributes between occupied and unoccupied logs.
We used the residuals of a linear regression of logtransformed lizard mass vs. log-transformed SVL
as a metric of lizard condition. SVL is primarily an
indicator of lizard age, and greater mass to SVL
ratio indicates a more robust lizard for its length.
We tested for correlations among log, soil, and
lizard attributes.
RESULTS
We found tropical night lizards under fewer than
half of all the logs that we overturned. Not all logs
that we initially overturned were considered
structurally viable. However, we easily found
Table 1. Two-tailed t-test statistics showed that
occupied and unoccupied logs did not exhibit
differences in any of the log and soil attributes.
occupied and unoccupied structurally viable logs.
We considered logs in contact with one another to
be one log. Tropical night lizards were often found
under flat, board-like logs. When we overturned
logs, lizards generally remained still and were
easily caught, only occasionally biting the handler.
Up to three tropical night lizards were found
under a single log. Log occupation was not related
to percent soil moisture, organic horizon depth,
log decomposition, soil temperature, log
temperature, area under log, or log volume (Table
1). We found other organisms under logs including
centipedes, millipedes, beetles, grubs, worms, a
land crab, and a caecilian (Fig. 2). However, the
presence of other organisms was not related to log
occupation by lizards (Χ2 = 0.18, P = 0.67, df =
1).
Along the trail versus in the arboretum,
organic horizon depth was greater (t = 2.1, P =
0.05, df = 20) and logs were more decomposed (t
= 3.21, P = 0.004, df = 24). Lizard condition
tended to be more robust in the arboretum than
along the trail (t = 2.21, P = 0.058, df = 8), but no
Table 2. Correlation matrix for relationships between log, soil, and lizard attributes. Lizard condition decreased as
both soil and log temperature increased. No other log and soil attributes were significantly correlated with lizard
condition.
96
Dartmouth Studies in Tropical Ecology 2012
A
Lizard condition
0.150
B
0.150
0.100
0.100
0.050
0.050
0.000
0.000
-0.050
-0.050
-0.100
-0.100
23
23.5
24
24.5
25
Soil temperature (°C)
23
24
25
26
27
Log temperature (°C)
Figure 1. Lizard condition decreased with (A) soil temperature (slope = -0.032 ± 0.014, P = 0.069, r2
= 0.35) and (B) log temperature (slope = 0.070 ± 0.033, P = 0.055, r2 = 0.39).
other attributes varied between locations. Lizard
condition was negatively correlated with both soil
temperature and log temperature (Table 1, Fig. 1).
No other log attributes predicted the robustness of
lizard condition (Table 2), nor did the presence of
other organisms (t = 0.06, P = 0.96, df = 2).
DISCUSSION
Safety from predators and access to prey can affect
the success of ambush predators. Our data suggest
that tropical night lizards use logs as shelters but
not as hunting grounds. Tropical night lizards were
not more likely to be found under logs with
attributes promoting arthropod abundance, nor
were the lizards in better condition under these
logs (Table 1). Tropical night lizards appear to
select logs irrespective of the quality of hunting
under the log. This suggests that the lizards are
choosing logs mainly just for shelter. Since over
half of structurally viable logs surveyed were
unoccupied and resources did not differ between
occupied and unoccupied logs, logs do not seem to
be a limiting resource for tropical night lizard
populations.
We found near-significant negative
correlations between lizard condition and both log
temperature and soil temperature (Fig. 1). Since
lizards do not appear to be foraging under logs, it
is possible that they maintain a lowered metabolic
rate by resting under cooler logs. Since higher
resting body temperatures will lead to more rapid
utilization of fat stores, keeping cool may be
energetically advantageous. Many animals reduce
their metabolic rates when they are not active to
prevent this unnecessary burning of fat stores
(Geiser 2004).
If tropical night lizards reduce their metabolic
rate during the day and utilize logs only as hiding
places, this may indicate that tropical night lizards
are nocturnal, hiding under logs during the day
and leaving to forage at night. Tropical night
lizards have been known to return to the same log
each day for shelter (Brian Folt, pers. comm.), and
our observations showed that several of the
tropical night lizards found on the first day of
study remained under the same log the next day.
Log preference may explain why tropical night
lizards are returning to the same logs each day to
rest. Such preferences may be due to log quality
variations influenced by environmental conditions
that our study did not account for.
Logs appear to primarily provide shelter for
and satisfy the physiological needs of tropical
night lizards at rest. The preferences of tropical
night lizards might help us understand factors that
influence shelter selection among lizards and other
ectotherms. Physiological requirements and
constraints, along with predator evasion, may
drive the spatial choices that animals make
throughout their lives. An understanding of
conditions that determine habitat suitability may
provide insight into how an animal uses that
habitat.
97
La Selva
ACKNOWLEDGEMENTS
Special thanks to Matthew P. Ayres of Dartmouth
College for helping us analyze our data and Bryan
Folt and providing information on L.
flavimaculatum.
LITERATURE CITED
Cooper, W. E. 1994a. Chemical discrimination by
tongue-flicking in lizards: a review with
hypotheses on its origin and its ecological and
phylogenetic relationships. Journal of
Chemical Ecology 20: 439-487.
Folt, Brian. Personal communication, February
2014.
98
AUTHOR CONTRIBUTIONS
All authors contributed equally.
Geiser, F. 2004. Metabolic rate and body
temperature reduction during hibernation and
daily torpor. Annual Review of Physiology 66:
239-274.
Savage, J. M. 2002. The amphibians and reptiles
of Costa Rica: a herpetofauna between two
continents, between two seas. University of
Chicago Press, Chicago. 498-499.
Dartmouth Studies in Tropical Ecology 2014
BEHAVIORAL THERMOREGULATION IN FACULTATIVELY POIKILOTHERMIC SLOTHS
(BRADYPUS VARIEGATUS AND CHOLOEPUS HOFFMANNI)
EMILY R. GOODWIN
Faculty Editor: Matthew P. Ayres
Abstract: Behavioral thermoregulation is critical to maintenance of body temperature for many organisms. Both
homeotherms and poikilotherms use this strategy- homeotherms use it to decrease metabolic costs associated with
maintenance of a constant body temperatures and poikilotherms use it to alter their body temperature. As facultative
poikilotherms, sloths are unusual among mammals in that they allow their body temperature to change with ambient
temperature. This strategy minimizes metabolic energy expenditures and is important for sloths as they eat lowquality folivorous diets that are digested slowly by symbiotic microbial fermentation. When sloth body temperature
is high, their metabolic rate and fermentation rates increase. I hypothesized that sloths behaviorally thermoregulate
to maximize sunlight cover throughout the day through both moving throughout the canopy and sleeping in postures
that maximize exposure of their bodies to direct sunlight. I focally observed six sloths of two species – one
Bradypus variegatus individual and five Choloepus hoffmanni individuals over three days. Sun coverage was
unrelated to sleep posture or canopy height, but sloths occasionally performed energetically expensive movements
both vertically and horizontally to canopy gaps where sun exposure was higher. This suggests that the
thermoregulatory and digestive benefits of sloth basking behavior outweigh the energetic costs of both horizontal
and vertical movements.
Key words: behavioral thermoregulation, Bradypus variegatus, Choloepus hoffmanni, poikilothermy, sleep posture
INTRODUCTION
Temperature influences many biological and
physiological processes in organisms across taxa.
Homeotherms maintain a constant body
temperature through metabolic processes, while
the body temperature of poikilotherms conforms to
environmental temperature, resulting in lower
energetic requirements. In both homeotherms and
poikilotherms, behavioral adjustments such as
basking and group huddling can decrease the
energetic costs of thermoregulation (Terrien et al.
2011). Homeothermic organisms such as birds and
mammals may congregate in close groups to
maintain their body temperatures, while
poikilothermic organisms such as reptiles may
bask in the sun to increase their body
temperatures, and move into shaded locations to
lower their body temperatures.
While most endotherms are homeotherms and
most ectotherms are poikilotherms, the
thermoregulatory strategies of organisms can vary.
Sloths are endothermic mammals that utilize
facultative poikilothermy, allowing their body
temperature to conform to ambient temperature,
unlike most other endotherms (Britton & Atkinson
1938). Sloths are specialized folivores with multicompartment stomachs for symbiotic microbial
fermentation of fiber. Due to their extremely low
basal metabolic rate (BMR), they must devote a
large proportion of their energy budget to
digestion (Goffart 1971). Thus, poikilothermy
allows sloths to minimize non-digestive energetic
expenditures (Costello and Rieb 2012). At high
ambient temperatures, sloths are able to decrease
energy spent on thermoregulation and increase
digestive rates. The evolution of poikilothermy in
sloths may be a strategy for aiding in digestion of
fiber-rich leaves while minimizing energetic costs.
If high temperature increases the efficiency of
sloth digestion, sloths would benefit from
maximizing their body temperatures through
basking. Sunning behavior has also been proposed
to enhance the activity of symbiotic gut bacteria
(Goffart 1971, Urbani and Bosque 2007).
Increased sun exposure can be achieved in two
ways: either by physically moving throughout the
canopy to sun gaps, or by sleeping in postures that
will maximize the surface area of their bodies
exposed to the sun. However, low canopy
exposure increases predation risk from visual
predators such as the harpy eagle (Harpia
harpyja), a major sloth predator that plucks sloths
directly from the canopy (Touchton et al. 2002).
Thus there is a tradeoff between the physiological
99
La Selva
benefits of sunning behavior and increased
energetic expenditure to seek light gaps.
I investigated whether behavioral
thermoregulation occurs in the brown-throated
sloth (Bradypus variegatus) and Hoffman’s twotoed sloth (Choloepus hoffmanni) in Heredia,
Costa Rica. Given the tradeoff between enhanced
digestion and predation risks, I hypothesized that
sloths would change their location in the canopy
throughout the day with temperature. Sloths could
move into sunnier locations in the mornings and
evenings, when ambient temperatures are lower
and the costs of increased digestive rates outweigh
the risk of predation. Conversely, when ambient
temperatures are higher during midday,
descending from the canopy for protection from
visual predators would bear minimal costs to
digestion. However, since moving throughout the
canopy is energetically expensive for sloths, they
may remain in low exposure areas and instead
adjust their sleeping posture within a tree
throughout the day to maximize sun exposure.
METHODS
I observed six sloths, five C. hoffmanni and one B.
variegatus, at La Selva Biological Station, Costa
Rica, between the hours of 0700 and 1745 on 1517 February 2014. All sloths were found within a
1 km stretch of secondary rainforest. Two C.
hoffmanni individuals, one with an infant, were
found in close proximity to one another, and were
observed resting while cuddling. One B.
variegates also had an infant. I used focal
sampling (Altmann 1974) on individuals for 10
minute periods approximately every hour, using
binoculars (8x magnification) and a spotting scope
(20x magnification) to observe sloth posture and
behavior.
I recorded proportion of canopy cover,
proportion of direct sun cover on body, sloth
posture, sloth behavior, and weather conditions for
each focal period (Table 1). I assessed sloth height
in the canopy using a hypsometer. Ambient air
temperature was collected at three locations placed
approximately 2 m off the ground under focal
sloth trees using a HOBO logger recording every
10 minutes.
100
Statistical Analyses and Energy Budget
Calculations
I calculated the daily activity budget of the
sloths by averaging the amount of time spent in
any given position, canopy cover, sun cover, sloth
behavior, and postures for each focal sample. I
used chi-square tests to analyze the effects of sun
cover and canopy cover on sloth posture. I
calculated the energy expenditure of vertical
climbing and horizontal movements using
published regression models for mammals
Table 1. Variables assessed during sloth focal sampling.
Variable
Description
Canopy Cover
Low
0-30% of sloth’s body covered by
leaves
Medium
30-60% of sloth’s body covered by
leaves
High
60-100% of a sloth’s body covered by
leaves
Sun Cover
Low
Medium
High
Sloth Behavior
0-30% of sloth’s body being hit by direct
sunlight
30-60% of sloth’s body being hit by
direct sunlight
60-100% of sloth’s body being hit by
direct sunlight
Resting
Stationary, eyes closed
Traveling
Horizontal or vertical locomotion, either
hanging or upright
Scanning
Eyes open, moving head around
scanning its environment
Feeding
Biting or chewing leaves, reaching for
branches
Sloth Posture
Huddle sitting
Stationary, sitting on at least one
branch, getting support from additional
branches.
Huddle
hanging
Stationary, holding self up with limbs by
holding a branch or tree trunk.
Extended
laying
Stationary, reclined with back on a
branch, ventral-side up, often
supporting itself with its forelimbs.
Dartmouth Studies in Tropical Ecology 2014
(Equation 1 and 3) (Pontzer 2011). All statistical
analyses were performed using JMP 11.0 software
(SAS Institute, Cary, NC).
I used the following equation to estimate the
energetic cost of climbing movements:
(Eq. 1)
afternoon, a favorite food tree of many foliviorous
mammals due to the high nutrient and low fiber
content of leaves (Urbani and Bosque 2007).
There was no effect of time of day or temperature
on sloth movement on height in the canopy.
A
B
C
D
In the above equation, mass = average weight
for each species, from literature (4.25 kg for B.
variegatus, 5.5 kg for C. hoffmani), g= 9.8 m/s2, h
= vertical distance moved (in meters). I calculated
energetic efficiency from regressions derived from
studies of primate energetics (Pontzer 2011):
(2)
It should be noted that efficiency is essentially
invariant with body mass.
For horizontal movements, I estimated cost of
transport from studies of energetics in a large
sample of mammals (Pontzer 2011):
(3)
Here, mass = average weight for each species from
literature. I then multiplied cost of transport by
horizontal distance traveled to acquire the energy
required to make the movements (in Joules).
Although these models derived from other
mammals do not reflect the idiosyncrasies of sloth
physiology and energetics, they provide useful
estimates for ranging costs, which have not been
directly measured or estimated in previous studies.
RESULTS
Daily temperatures ranged from 22°C to 31°C
during the three-day study period. Temperature
peaked around noon and was lowest at dawn. The
data were consistent across the three HOBO
loggers. Weather was consistently sunny
throughout the three sampling days.
Sloths were found in a total of seven trees at
positions high in the canopy. Sloths spent 90% of
their daily activity budget resting, 6.5% traveling
to new locations, and 3.2% scanning (Table 1). I
never observed sloths feeding during the daytime,
which was consistent with the reported nocturnal
nature of sloths (Goffart 1971). However, one
sloth moved to a Cecropia tree in the late
Figure 1. (A) C. hoffmanni in the extended laying resting
posture. (B) B. variegatus in the huddle-sitting resting
posture. (C) B. variegatus in the huddle-hanging resting
posture. Note the infant holding on to her ventral side.
(D) C. hoffmanni in the hanging traveling posture. All
photos courtesy of the author.
When resting, sloths spent 47% of their time
in extended laying positions, 29% of their time in
huddle-sitting positions, and 22% of their time in
huddle-hanging positions (Fig. 1). There was no
effect of time of day or temperature on sloth
posture. Posture did not vary with sun exposure
(chi-square = 7.28, P = 0.12, df = 4; Fig. 2), but
did vary with degree of canopy cover (chi-square
= 26.3, p <0.0001, df = 4; Fig. 3). On the
afternoon of 16 February, it rained (while still
sunny) for approximately twenty minutes. During
the rainfall, the focal sloth (C. hoffmanni) changed
postures from extended laying to huddle-sitting
until the rainfall stopped.
When repositioning or moving within the
canopy, sloths most often repositioned to another
spot with the same sun coverage (85%), followed
by repositioning to a spot with higher sun
coverage (12%) (chi-sqaure = 65.59, P = 0.000, df
= 8; Fig. 4). Sloths were rarely seen moving to a
101
La Selva
location with less sun coverage than the previous
position (3%; Fig. 4). When traveling, sloths spent
89% of time hanging (Fig. 1) and 11% of time
upright.
100%
90%
% Time in Posture
80%
70%
huddlesitting
60%
50%
huddlehanging
40%
extended
laying
30%
20%
10%
0%
low
medium
high
Sun Exposure
Figure 2. Sloth resting posture did not vary with sun
exposure (chi-square = 7.28, P = 0.12, df = 4).
100%
90%
% Time in Posture
80%
70%
huddlesitting
60%
huddlehanging
50%
40%
extended
laying
30%
20%
10%
0%
low
medium
high
Canopy Cover
Figure 3. Sloth resting posture varied with canopy cover
(chi-square = 26.3, p <0.0001, df = 4).
Across observation periods, five large (> 0.5
m) horizontal and vertical movements were
observed. Each movement resulted in a new
position in a canopy gap with equal or higher sun
cover (Table 2). Estimated energy expenditures for
these movements were highest for sloths moving
from low to high sun exposure (Table 2).
102
DISCUSSION
As facultative poikilotherms, sloths use ambient
temperature changes to increase their body
temperature and thus digestive rate. However,
sloths did not try to get higher into the canopy
during the cooler parts of the day, or descend into
the canopy at high temperatures to avoid the
temperature-predation risk tradeoff. It may be that
predation risk varies little with position in canopy.
Moreover, postural and morphological
mechanisms could be more relevant for avoiding
predation in sloths, since they spend most of the
day sleeping. For example, sloths have specially
adapted fur to encourage algae growth on the
pelage, as well as very slow body movements,
which may help reduce visual detection from
predators. Additionally, I saw that sloths in highexposure canopy gaps were more likely to adopt
huddling postures that minimized body surface
area, possibly making them more cryptic than they
would be in the extended laying posture.
When a sloth repositioned, it almost always
led to a new position in equal or higher sun
coverage (Fig. 4). Additionally, all large
horizontal and vertical movements ended with
sloths in canopy gaps with equivalent or higher
sun cover (Table 2). These results suggest that the
maximization of sun cover on body surface area
from any given resting posture in the canopy is
less beneficial than relocating to a new canopy gap
with higher exposure, possibly because of changes
in the position of the sun throughout the day.
Although the range of movements observed in this
study result in low energy expenditures per
movement, the costs of these movements are more
expensive for sloths as compared to other
mammals of similar body size due to their low
basal metabolic rate (147 kJ/(kg*day); (Nagy and
Montgomery 1980). Both of the highest observed
sloth energetic expenditures were associated with
a move from low to high sun cover (Table 2).
Relocations that result in increased sun exposure
appears to be worthwhile despite the costs,
perhaps because of an increased digestive rate
associated with higher body temperature. I
therefore suggest that the benefits of increased
digestive efficiency outweigh the energetic costs
of canopy movements.
Dartmouth Studies in Tropical Ecology 2014
Table 2. All long distance movements during or between focals (defined >0.5 m in distance). Energy expenditures
were calculated using Equation 1 for vertical climbing & Equation 2 for horizontal movement.
Date
Time
Species
Movement
Change in Sun
Exposure
Observations
Energy
Expenditure (J)
2/15/14
8:45
B. variegatus
vertical descent 0.7 m
low to medium
Descent partially out of canopy and
into bare patch near tree trunk
N/A
2/15/14
10:59
C. hoffmanni
vertical ascent 0.9 m
high to high
Partial emerging from canopy; head
above the leaves
31.52
2/16/14
9:01
C. hoffmanni
vertical ascent 1.3 m
medium to high
Movement towards canopy gap
45.53
2/16/14
15:39
C. hoffmanni
horizontal movement
22.3 m
low to high
Between focals, movement to large
canopy gap one tree over
760.6
2/16/14
15:51
B. variegatus
vertical descent 4.3
m, ascent 5.0 m
low to high
Between focals, left current tree
and then moved to the canopy of a
mostly bare Cecropia
Descent: N/A
time to maintain their body temperatures.
Associated increases in body temperature may aid
in fermentation and digestion of their folivorous
diets by symbiotic microbial communities.
0.7
58%
0.6
0.5
Probability
Sloths spent the highest proportion of their
resting in the extended laying posture, followed by
the huddle-sitting posture, and the huddle-hanging
posture. Both of the huddle postures are
predominately upright positions. Thus sloths spent
approximately half of their time upright and the
other half reclining, despite the reclining posture
exposing their ventral sides to the sun and
potentially aiding in digestion more so than the
other postures. Alternatively, the upright posture
of could also be important in digestion. In
ruminant herbivorous mammals, upright posture
had been hypothesized to be an adaptation to
increase stratification of forestomach contents,
allowing larger food particles to pass more quickly
(Clauss 2004). The strictly foliviorous B.
variegatus spent a higher proportion of time in the
two upright postures (91%) than C. hoffmanni
(22.2%), which consume more readily fermentable
fruits, flowers, and buds in its diet (Meritt 1985).
C. hoffmanni’s more easily fermentable diet may
increase the freedom of individuals to vary resting
positions, while B. variegatus must rest upright
and therefore ease digestion. This is consistent
with the forestomach stratification hypothesis
(Clauss 2004).
Sloths did appear to utilize behavioral
thermoregulation as a function of their facultative
poikilothermy. Sloths relocated throughout the day
to get to canopy gaps, despite the high energy
costs of those movements. This sunning behavior,
in addition to dietary knowledge of both species of
sloth and its effects on posturing, indicate that
sloths must maximize their use of their resting
Ascent: 139.16
0.4
0.3
19%
0.2
0.1
0
4%
8%
12%
ML
LL
MH
MM
HH
Change in Sun Exposure after
Repositioning
Figure 4. Most sloth repositioning led to maintenance of
sun exposure (85%), followed by repositioning or new
postures that led to a higher percntage of sun exposure
(18%). Rarely did sloths reposition to a posture with less
sun exposure (4%) (chi-sqaure = 65.59, P = 0.000, df =
8.). Notation (e.g. ML) represents sun exposure before
reposition (M) and sun exposure after reposition (L)
where L= low (0-30%), M= medium (30-60%), and H=
high (60-100%).
103
La Selva
ACKNOWLEDGEMENTS
Thanks to the guides of La Selva Biological
Station for their help and humor in locating sloths.
I am also indebted to V. Venkataraman and J.
Trout-Haney for all of their assistance and support
at all steps of this process.
LITERATURE CITED
Altmann, J. 1974. Observational study of
behavior: sampling methods. Behavior 49:
227- 267.
Britton, S.W. and W.E. Atkinson. 1938.
Poikilothermism in the sloth. Journal of
Mammalogy 19: 48-51.
Clauss, M. 2004. The potential interplay of
posture, digestive anatomy, density of ingesta
and gravity in mammalian herbivores: why
sloths do not rest upside down. Mammal
Review 34: 241-245.
Costello, R., and J. Rieb. 2012. How Choloepus
hoffmanni’s poikilothermy preserves the sloth
niche. Dartmouth Studies in Tropical Ecology
2012, pp. 118-123.
Goffart, M. 1971. Function and form in the sloth.
Oxford: Pergamon Press.
Merritt D.A. 1985 The two-toed Hoffman’s sloth,
Choloepus hoffmanni. The evolution and
ecology of armadillos, sloths and
vermilinguas. Washington, D.C.: Smithsonian
Institution Press pp 333-41.
Nagy, K.A., and G.G. Montgomery. 1980. Field
metabolic rate, water flux, and food
consumption in the three toed sloth (Bradypus
variegatus). Journal of Mammalogy 61: 465472.
Pontzer, H. 2011. From treadmill to tropics:
calculating ranging cost in chimpanzees.
Primate Locomotion (eds. D'Aout K, Vereecke
EE). Springer.
Terrien, J., M. Perret, and F. Aujard. 2011.
Behavioral thermoregulation in mammals: a
review. Frontiers in Bioscience 16: 1428.
Touchton, J.M., Y. Hsu, and A. Palleroni. 2002.
Foraging ecology of reintroduced captive-bred
subadult harpy eagles (Harpia harpiya) on
Barro Colorado Island, Panama. Ornitologia
Neotropical (The Neotropical Orinthological
Society) 13.
Urbani, B. and C. Bosque. 2007. Feeding ecology
and postural behavior of the three-toed sloth
(Bradypus variegatus flaccidus) in northern
Venezuela. Mammalian Biology 72: 321-329.
104
Dartmouth Studies in Tropical Ecology 2014
PATTERNS IN CORAL ASSEMBLAGES OF A PATCH REEF SYSTEM
MAYA H. DEGROOTE, EMILY R. GOODWIN, AARON J. MONDSHINE, AND ELLEN R. PLANE
Faculty Editor: Celia Y. Chen
Abstract: Bleaching, disease, and natural disasters are leading to increased destruction and degradation of coral reef
habitats. Coral habitat loss reduces overall reef size and continuity, increasing the prevalence of patchy reef
environments. In order to determine what affects recruitment of corals in patchy environments, we conducted a
study of coral communities on reef patches in Grape Tree Bay, Little Cayman. Coral recruitment may follow the
theory of island biogeography, which predicts that larger islands and islands closer to the source population, in this
case the main reef, will have greater species richness. We found that larger reef patches had greater species richness,
diversity, and percent coral cover than smaller reef patches. However, we found no significant effect of distance
from the main reef on coral assemblages, indicating that distance is not a limiting factor for coral dispersal and
recruitment in this system. We conclude that conservation efforts should focus on preserving areas with larger reef
patches, since these patches foster greater species diversity and tend to have higher coral cover.
Key words: coral communities, island biogeography, patch reef, SLOSS, species diversity
INTRODUCTION
Globally, coral reef health is becoming
increasingly threatened due to habitat destruction
and degradation (Nyström et al. 2000, Hughes et
al. 2003). Many factors can lead to coral reef
habitat destruction, including disease, coral
bleaching, and natural disasters, all of which are
exacerbated by global climate change (Shepard et
al. 2009). Reef habitat loss decreases overall reef
size and continuity, reducing the resilience of reefs
to subsequent disturbance (Nyström et al. 2000).
The resulting patch reefs, or small coral reefs that
are isolated from a larger, continuous main reef,
therefore may become more common. Changes in
species richness, diversity, and percent live coral
cover caused by habitat loss can impact coral
health and resilience (Hughes et al. 2003, Roff and
Mumby 2012), so it is important to understand
what affects coral recruitment in patchy reef
environments.
Recruitment of corals to patch reefs may be
comparable to the species colonization of islands,
as described by the theory of island biogeography
(MacArthur and Wilson 1967). Patch reefs are
islands of suitable habitats that are surrounded and
separated by areas of unsuitable habitats
(MacArthur and Wilson 1967). The theory
predicts that species diversity on an island
increases with island size because immigrants are
more likely to find and migrate to larger islands.
Species-area relationships show that larger areas
are correlated with increased species diversity
because of a variety of factors including increased
habitat heterogeneity and resource availability
(Shen et al. 2009). Species diversity also increases
with proximity of the island to the mainland, or
source population. Since immigrants must travel
shorter distances, a larger variety of species with
different mobility and dispersal strategies are able
to travel to the island. The theory of island
biogeography may help to explain patterns of coral
diversity, measured by species richness and
abundance, in a patch reef system.
Corals have a variety of reproductive
strategies that may influence their recruitment to
near and far reefs. Coral species can be
hermaphroditic (both sexes in one individual), or
gonochoristic (two distinct sexes), and many
Scleractinian corals can reproduce asexually
(Richmond & Hunter 1990). Location of
fertilization can also vary, either occurring within
the individual (brooding) or externally when
gametes are released in the water column
(broadcast spawning) (Richmond and Hunter
1990). In both cases, planktonic larvae must travel
through the water column and land on a suitable
substrate before a colony can form. Because of
this broadcast spawning strategy, recruitment may
not be limited by the same factors that limit
terrestrial organisms, and thus corals may not
follow the predictions of island biogeography
theory.
We conducted this study to determine whether
the distribution of corals in a patch reef system are
105
Little Cayman
Statistical Analyses
We conducted linear regressions to test whether
water depth influenced either coral species
richness or percent coral cover. We performed a
linear regression of reef patch surface area and
distance from the reef crest and found that these
factors were not correlated, meaning that our data
included reef patches of all sizes across a range of
distances from the reef crest. We calculated
106
Species richness
12
10
8
6
4
2
0
0
50
100
0
50
100
3
Species diversity
METHODS
We surveyed 30 reef patches between the main
reef and shore in Grape Tree Bay, Little Cayman,
on 26-27 February 2014. A reef patch was defined
as a patch of reef separated by at least two meters
from the fringing reef crest and from all other reef
patches. We haphazardly selected reef patches
across a range of both sizes and distances from the
main reef, defined as the reef crest. We calculated
reef patch surface area as an idealized cylinder,
excluding the base, by measuring circumference
and height of each patch using measuring tape. In
addition, at each reef patch we measured water
depth and distance from the reef crest on an axis
perpendicular to the shoreline.
We performed a visual census of live coral
communities at each reef patch, identifying all
species and counting the abundance of individual
colonies of each species. Three observers
estimated the percentage of the reef patch that was
covered by live coral. We averaged these estimates
to yield a measure of percent coral cover.
species diversity for each reef patch using the
Shannon-Weiner Diversity Index. Since speciesarea relationships are generally described by
logarithmic curves, we then fit logarithmic
regressions comparing patch surface area to
species richness, species diversity, and percent
coral cover. We then performed linear regressions
comparing distance from the reef crest to species
richness, species diversity, and percent coral
cover. For coral species that were present on at
least five reef patches, we fit regressions of
abundance to reef patch surface area and distance
from the reef crest. We categorized species as
stony (Scleractinia) or soft (Alcyonacea and
Gorgonacea) and tested for a relationship between
their abundances using a linear regression.
2.5
2
1.5
1
0.5
0
Percent coral cover
consistent with the predictions of island
biogeography. If so, coral species richness should
increase with (1) reef patch size, and (2) proximity
of the reef patch to the main reef. Larger reef
patches should have greater coral species richness
since they have greater surface area and greater
habitat heterogeneity. While many previous island
biogeography studies have measured species
diversity using species richness, we also predict
that patterns of overall species diversity, and
percent coral cover will follow these trends. It is
also possible that coral dispersal and recruitment
are not limited by distance from the source
population, so there will be no effect of proximity
to the reef crest on species richness. In this case,
other factors, including water depth, tides, current,
and dispersal strategy, may determine coral
recruitment patterns in patch reefs.
70
60
50
40
30
20
10
0
0
50
100
Reef patch surface area (m2)
Figure 1. Species richness, species diversity, and
percent coral cover increased logarithmically with reef
patch surface area.
Dartmouth Studies in Tropical Ecology 2014
Table 1. Species richness, species diversity, and
percent coral cover were described by logarithmic
relationships with reef patch surface area (x).
Equation
Species
richness
Species
diversity
Percent
coral cover
P
r2
y = 2.42 + 1.55(log(x))
<0.0001
0.47
y = 1.33 + 0.28(log(x))
0.0053
0.25
y = 9.91 + 11.98(log(x))
<0.0001
0.52
DISCUSSION
The patterns of coral species composition in reef
patches were consistent with the species-area
relationships predicted by island biogeography
theory. We found that species richness increased
with reef patch surface area (Fig. 1). However, we
found no relationship between distance to reef
crest and species richness (Fig. 2), indicating that
dispersal mechanisms from the source reef
population may differ from those operating in
other island scenarios, especially in terrestrial
systems. Species diversity and percent coral cover
followed similar trends for both surface area of
reef patches and proximity to reef crest. Although
we counted abundance of each coral species by
counting number of individual colonies, whether
or not each colony was a distinct colonization
event was unclear.
Species richness
12
10
8
6
4
2
0
0
10
20
30
40
Species diversity
3
2
1
0
0
Percent coral cover
RESULTS
We identified 20 total coral, hydrozoan, and
zoanthid species (Appendix 1). The most common
species (stony corals: Agaricia agaricites, Porites
astreoides; soft corals: Briareum asbestinum and
Eunicea spp.), were found on over half of reef
patches surveyed.
Water depth ranged from 1.3 to 2 m and did
not have an effect on species richness, species
diversity, or percent coral cover. Reef patch
surface area ranged from 1.5 to 69.6 m2, and
distances between reef patches and the fringe reef
crest ranged from 3.8 to 68.4 m.
Reef patches with greater surface area had
higher species richness, species diversity, and
percent coral cover (Table 1, Fig. 1). Exclusion of
the reef patch with a surface area of 69.6 m2 did
not have an effect on any of the three logarithmic
relationships. Larger patches had more colonies of
B. asbestinum (P < 0.0001, r2 = 0.73) and
Montastrea annularis (P < 0.0001, r2 = 0.49).
Distance from the reef crest did not have an
effect on species richness, species diversity, or
percent coral cover (Fig. 2). Isophyllastrea rigida
was the only species to demonstrate a distance
effect, and was found in patches at least 40 m from
the reef crest. There was no difference between
stony and soft coral abundance in response to any
factor. Patches with more stony corals also had
more soft corals (P = 0.01, r2 = 0.20).
70
60
50
40
30
20
10
0
0
20
40
60
80
50
100
Distance from reef crest (m)
Figure 2. Distance from the reef crest did not have an
effect on species richness, species diversity, or percent
coral cover on reef patches.
The theory of island biogeography states that
larger islands are likely to contain more niches
(MacArthur and Wilson 1967). With increased
habitat heterogeneity, immigrating species are
more likely to find an available niche to occupy
and therefore be more successful. Our findings,
which show that reef patches with more surface
area had greater species richness, could be the
result of variations in physical habitat
heterogeneity. However, additional abiotic and
107
Little Cayman
biotic factors may also be influencing niche
availability. Further studies should investigate how
island size affects coral recruitment and
colonization by investigating the abundance of
new coral colonies on both large and small reef
patches.
Larger reef patches have previously been
associated with higher fish abundance and species
richness (Lobben and Cheek 2008, Chowdhury et
al. 2012). Increased algal herbivory and coral
predation by fish could help explain why we found
higher coral species richness on larger reef
patches. Algal herbivory reduces algal-coral
competition, creating space for coral recruitment
and increased abundance, while intermediate
levels of coral predation could allow for increased
species diversity.
We found no relationship between species
richness and distance from the reef, indicating that
distance may not be a limiting factor for coral
dispersal. Additionally, neighboring reef patches
may affect species richness and therefore the reef
crest may not be the only source from which coral
species emigrate. In this study, patch reefs were no
more than 70 m from main reefs. Reef patches
located at greater distances from a reef crest may
exhibit coral distributions consistent with the
theory of island biogeography because dispersal
may be a limiting factor at larger spatial scales.
A better understanding of coral recruitment to
patchy reef environments may help determine
appropriate sizes and locations for marine
protected areas. Our findings also have
implications for the SLOSS (Single Large or
Several Small) debate in ecology, which explores
whether it is preferable to preserve a single large
area or several smaller areas. The theory of island
biogeography is addressed by the SLOSS debate
that uses diversity as an indicator for conservation
(Diamond 1975, Simberloff and Abele 1982). Our
study suggests that coral reef restoration efforts,
including implementation of artificial reefs, should
focus on larger reef patches, which promote
greater species richness. Conservation efforts
should concentrate on preserving areas with larger
reef patches, as these patches foster greater species
diversity and tend to have more coral cover.
LITERATURE CITED
Chowdhury, M.H.R., S.F. Flanagan, and N.B.
Frankel. 2012. Factors affecting fish diversity
on coral reef “islands.” Dartmouth Studies in
Tropical Ecology 2012, pp. 154-158.
Diamond, J.M. 1975. The island dilemma: lessons
of modern biogeographic studies for the
design of natural reserves. Biological
Conservation 7: 129–146.
Hughes, T.P., A.H. Baird, D.R. Bellwood, M.
Card, Connolly, S.R., Folke, C., Grosberg, R.,
Hoegh-Guldberg, O., Jackson, J.B.C.,
Kleypas, J., Lough, J.M., Marshal, P.,
Nystrom, M., Palumbi, S.R., Pandolfi, J.M.,
Rosen, B., and J. Roughgarden. 2003. Climate
change, human impacts, and the resilience of
coral reefs. Science 301: 930-933.
Lobben, T.J. and L.M. Cheek. 2008. Coral patches
on a back reef do not conform to diversitydistance predictions of island biogeography
theory. Dartmouth Studies in Tropical
Ecology 2008, pp. 161-164.
MacArthur, R. H., and E. O. Wilson. 1967. The
theory of island biogeography. Princeton
University Press, Princeton, NJ.
Nyström, M., C. Folke and F. Moberg. 2000.
Coral reef disturbance and resilience in a
human-dominated environment. Trends in
Ecology and Evolution 15: 413-17.
Richmond, R.H., and C.L. Hunter. 1990.
Reproduction and recruitment of corals:
comparisons among the Caribbean, the
Tropical Pacific, and the Red Sea. Marine
Ecology Progress Series 60: 185-203.
Roff, G., and P.J. Mumby. 2012. Global disparity
in the resilience of coral reefs. Trends in
Ecology and Evolution 27: 404-413.
Shen, G., M. Yu, X. Hu, X. Mi, H. Ren, I. Sun,
and K. Ma. 2009. Species–area relationships
explained by the joint effects of dispersal
limitation and habitat heterogeneity. Ecology
90: 3033–3041.
Sheppard, C.R.C, S.K. Davey, and G.M. Pilling.
108
ACKNOWLEDGEMENTS
We would like to thank Katie Lohr of CCMI for
her assistance with coral identification.
AUTHOR CONTRIBUTIONS
All authors contributed equally.
Dartmouth Studies in Tropical Ecology 2014
2009. The biology of coral reefs. Oxford
University Press, Oxford, UK.
Simberloff, D. S. and L. G. Abele. 1982.
Refuge design and island biogeograpic
theory - effects of fragmentation.
American Naturalist 120: 41-56.
109
Little Cayman
GRAZING AND EPIBIONTS IN MARINE TURTLE GRASS BEDS
ALLEGRA M. CONDIOTTE, NATHANAEL J. FRIDAY, LARS-OLAF HÖGER,
AND ZACHARY T. WOOD
Faculty Editor: Celia Y. Chen
Abstract: Turtle grass serves as habitat for small herbivores and various types of epibionts. We examined the effects
of depth and turtle grass density and distribution on herbivory and epibiont cover in the shallow coastal water of the
Caribbean. We found that large-scale herbivory (cropped tips) increased with grass density while epibiont cover
increased with water depth. These results suggest that herbivores that take large bites (crop tips) selectively graze
dense areas of turtle grass, which are more easily visible and may provide protection from large predators, and that
epibiont herbivores may be limited to shallow water by predation.
Key words: community structure, epibionts, herbivory, habitat complexity, Thalassia testudinum, turtle grass
INTRODUCTION
Plants affect the physical structure of the
communities they inhabit. In marine ecosystems,
plants are not only a food source for herbivores
and decomposers; they provide shelter for
organisms, alter local currents, and are often a
substrate for other benthic organisms.
Turtle grass, Thalassia testudinum, is a
widespread submerged macrophyte in tropical
coastal marine systems commonly found
bordering the shoreline. As well as providing food
for herbivores, turtle grass creates a unique
substrate for communities of epibionts: organisms,
ranging from encrusting algae to hydrozoans, that
live fixed to the surface of the grass blades.
Photosynthetic epibionts, such as algae, can serve
as important primary producers in oligotrophic
systems, fixing carbon for higher trophic levels in
local communities (Kitting et al. 1984). Nutrientrich epibionts can make plants more attractive to
herbivores, increasing herbivore grazing on
epiphyte covered plants (Hay and Whal 1995).
However, as turtle grass provides the growth
substrate for these epibionts, its physical location
and patterning should impact epibiont distribution
and abundance. Herbivore activity should also be
affected by turtle grass distribution.
We examined how turtle grass structure
influences herbivory and epibiont coverage in
shallow coastal water in the Caribbean. If
herbivores graze indiscriminately and always
consume the same number of blades per area,
there should be a negative relationship between
herbivory and density, because less dense areas
will experience a proportionally greater effect of
110
herbivory. In turn, if herbivores graze
proportionally to the density of grass, there should
be no relationship between herbivory and density.
Finally, if herbivores favor high density areas of
grass, there should be a positive relationship
between herbivory and density. Seagrass
distribution is primarily limited by light
availability, particularly in deep water (Ralph et al.
2007). If photosynthetic epibionts have similar
light limitations to their host macrophytes, they
should grow throughout turtle grass beds.
However, it is possible that photosynthetic
epibionts have more stringent light requirements
than seagrasses, in which case turtle grass may be
able to grow where epibionts cannot.
METHODS
Data collection
We studied turtle grass at the Little Cayman
Research Centre, Cayman Islands, on 27 February
2014. We surveyed six parallel 18-meter transects
perpendicular to the shore, starting two meters
from the shoreline. We spaced transects two
meters apart. At two-meter intervals along each
transect, we haphazardly selected five blades of
turtle grass, which we later assessed for herbivory
and epibiont cover in the lab. Epibiont cover was
assessed visually by estimating and averaging the
percent cover from each side of each blade. To
assess herbivory we counted the bites taken out of
the blades and noted whether the tip of the blade
had been cropped off. We defined our tips cropped
variable as the proportion of blades sampled at
each site that had cropped tips, and we defined our
bites variable as the average number of small
Dartmouth Studies in Tropical Ecology 2014
(<0.5 cm diameter) bites per blade at each site.
Tips cropped was classified as large scale
herbivory, and small bites were classified as small
scale herbivory, reflecting the size of animals most
likely capable of inflicting such damage. At each
site we calculated the density of turtle grass by
counting blades within a 25.5*25.5 cm quadrat,
and measured the average height of the grass and
depth of the seafloor.
Statistical Analyses
To examine how turtle grass bed characteristics
influenced herbivory and epibiont cover, we fit
several linear models with proportion of tips
cropped, bites per leaf, and epibiont cover as
dependent variables. For these dependent
variables, we separately tested all individual and
interaction effects.
Table 1. Model results. H = bed height, D = density,
P = depth.
Response
Epibiont
cover
Tips
cropped
Bites
Effects and
direction
P
↓H
0.0081
↓D
0.18
↑P
0.023
↓H, ↓D, ↑H*D
0.073
↓D, ↓P, ↑D*P
0.11
↑P, ↓H, ↑P*H
0.034
↑H
0.63
↑D
0.039
↑P
0.67
↑H, ↑D, ↑H*D
0.20
↓D, ↓P, ↑D*P
0.12
↓P, ↓H, ↑P*H
0.13
↑H
0.087
↑D
0.14
↓P
0.12
↑H, ↑D, ↓H*D
0.31
↓D, ↓P, ↑D*P
0.17
↓P, ↓H, ↑P*H
0.20
RESULTS
Variation in turtle grass density was great, with the
ocean floor nearly invisible in many parts of the
beds and almost barren in others. Grass height
ranged from 6.5 to 28.0 cm (Appendix figures).
Variation in epibiont cover was also large, with
some blades almost completely covered and others
nearly bare.
Deeper areas had more than twice the epibiont
cover of shallower areas (Figure 2, Table 3, F1,58 =
5.497, P = 0.023). Although turtle grass height
also negatively affected epibiont cover, this effect
became insignificant when depth was included in
the model. Grazers did not appear to seek
epibiont-rich leaves; neither tip cropping or bites
per leaf were related to epibiont cover (F1,58 =
1.453, P = 0.23; F1,58 = 1.609, P = 0.21;
respectively). Tip cropping was more frequent in
denser turtle grass beds (Figure 1, Table 2, F1,58 =
5.478, P = 0.023). Bites per leaf were not related
to turtle grass height, seafloor depth, or epibiont
cover (Table 1). For all other models, see Table 1.
Figure 1. Tips cropped increased with turtle grass
density. R2 = 0.09.
Table 2. Regression coefficients for tips cropped
on turtle grass density
Effect
Est. ± Std. Err.
P
Intercept
-0.078 ± 0.090
0.031
Density
0.00012 ± 0.00005
0.023
111
Little Cayman
Figure 2. Epibiont cover increased with depth.
R2 = 0.09.
Table 3. Regression coefficients for epibiont cover
on water depth
Effect
Est. ± Std. Err.
P
Intercept
12.78 ± 13.30
0.34
Depth
49.10 ± 20.94
< 0.0001
DISCUSSION
We examined how the abiotic and biotic structure
of turtle grass beds affects herbivory and epibiont
cover. Turtle grass height, grass density, water
depth, epibiont cover, and herbivory varied in
different areas of the turtle grass (Appendix
figures). Large-scale herbivory (cropped tips) was
highest in near-shore areas of dense grass (Figure
1), while epibiont cover increased with depth
(Figure 2). Small-scale herbivory (all bites < 0.5
cm in diameter) did not vary with any of our
independent variables. Epibiont cover and
herbivory were not related, indicating that
herbivory and epibiont cover are driven by
different factors.
Our results suggest that herbivores that take
large bites (crop tips) selectively graze dense areas
of turtle grass. Large herbivores, such as bucktooth parrotfish, are visual grazers (Ogden 1976).
112
Other large herbivores such as green sea turtles
return to the same grazing plots to feed
periodically (Thayer et al. 1984). Because of this,
dense beds of turtle grass may be more visible or
desirable to these herbivores. Large herbivore
preference for denser patches of grass may
regulate seagrass expansion by limiting further
growth in already dense areas. Small-scale
herbivory did not vary with depth, grass height, or
grass density, suggesting that small herbivores
such as small fish and sea urchins feed
sporadically and do not show a preference for
denser patches.
Increased epibiont coverage in deep water
may also be related to predation. Animals that feed
on epibionts, including small snails, fish and crabs,
may be limited to shallow areas due to increased
predation by large fish in deep zones. Deep water
provides more space for large predators, making
these areas more dangerous for epibiont grazers
than shallow water. This could allow the epibiont
community to expand more easily in deep water.
Epibiont cover could also increase with depth
because of a decrease in ultraviolet radiation,
which is harmful to algae and can limit its growth
in shallow water (Bischof et al. 1998).
In tropical marine systems, seagrass structure
and distribution affect herbivory and epibiont
coverage, which in turn have implications for the
success of the macrophyte community and the
overall system. Our results suggest that herbivory
in these systems is density-driven, and that the
range of epibionts is limited by depth of water.
The density and distribution of turtle grass affects
organisms at multiple trophic levels, suggesting a
large-scale impact on the community.
ACKNOWLEDGEMENTS
We thank Celia Y. Chen, Jessica Trout-Haney, and
Vivek Venkataraman for their guidance and
direction in the experimental design and writing
process.
AUTHOR CONTRIBUTIONS
All authors contributed equally to the data
collection and writing process.
Dartmouth Studies in Tropical Ecology 2014
LITERATURE CITED
Bischof, K.., Hanlet., D., and Wiencke, C. 1998.
UV-radiation can affect depth-zonation of
Antarctic macroalgae. Maine Biology 131:
597-605.
Hay, M.E., and Fenical, W. 1988. Marine plantherbivore interactions: the ecology of
chemical defense. Annual Review of Ecology
and Systematics 19: 111-145.
Hay, M.E., and Whal, M. 1995. Associational
resistance and shared doom: effects of
epibiosis on herbivory. Oecologica. 102: 329340
Jensen, K. 1977. Effect of epiphytes on eelgrass
photosynthesis. Aquatic Botany 3: 55-63.
Kitting, L.C., Fry, B., and Morgan, D.M. 1984.
Detection of inconspicuous epiphytic algae
supporting food webs in seagrass meadows.
Oecologica 62: 145-149.
Ogden, J.C. 1976. Some aspects of herbivore-plant
relationships on Caribbean reefs and seagrass
beds. Aquatic Botany 2: 103-116.
Ralph, P.J., Durako, M.J., Enriquez, S., Collier,
C.J., and Doblin, M.A. 2007. Impact of light
limitation on seagrasses. Journal of
Experimental Marine Biology and Ecology
350: 176-193.
Thayer, G.W., Bjorndal, K.A., Ogden, J.C.,
Williams, S.L., Zieman, J.C. 1984. Role of
larger herbivores in seagrass communities.
Estuaries. 7: 351-376.
113
Little Cayman
BURROW OCCUPATION AND MOVEMENT IN ECHINOMETRA LUCUNTER
ANN M. FAGAN, REBECCA C. NOVELLO, AND CLAIRE B. PENDERGRAST
Faculty Editor: Celia Y. Chen
Abstract: Animals often select or create shelters in their environments that reduce predation risk and vulnerability to
environmental stressors. However, animals must devote time and energy to retaining these shelters, and they may
choose to abandon shelters when protective benefits do not outweigh the costs of maintenance and defense against
conspecifics. Echinometra lucunter is a tropical rock-boring sea urchin that creates and occupies burrows that offer
protection against predation and wave action. E. lucunter often return to the same burrow after nocturnal foraging
and must sometimes fight conspecifics to retain possession of high-quality burrows. However, burrow availability
could influence patterns of burrow selection and defense in E. lucunter. We hypothesized that in environments
where burrow availability is high, urchins would occupy burrows whose depths are proportional to their body sizes
because these burrows offer the greatest protection. We also hypothesized that in such environments, E. lucunter
would move freely between burrows and be more likely to leave a burrow to forage if the burrow was not
proportional to their body size. We found that in tidal pools with high burrow availability, E. lucunter occupy
burrows with depths that are proportional to their body size and fail to display site fidelity. Urchins occupying
shallower burrows or burrows that are not proportional to their body size are more likely to leave them to forage at
night. Population density and burrow availability appear to influence movement and site fidelity in E. lucunter,
suggesting that urchin behavior and habitat choice reflect local environmental conditions.
Keywords: body size, Echinometra lucunter, resource availability, site fidelity, urchin
INTRODUCTION
Many animals construct or select shelters in their
environment that facilitate protection, food
acquisition, or reproductive success. Since the
maintenance and defense of such shelters can be
energetically costly, selecting a high quality site is
critical. In areas with low shelter availability,
animals often prioritize the defense of adequate
shelters over the potentially expensive pursuit of
higher quality shelters. Animals in such areas
often display site fidelity, returning to the same
shelter over time and defending it from
conspecifics (Switzer 1997). In contrast, when
shelter availability is high, competition for shelter
may be weaker. Animals in these environments
may instead devote energy to foraging or seeking
out higher quality shelters instead of defending
current sites. They may also be less likely to
display site fidelity when competition for shelter is
reduced, as the potential benefits of exchanging
sites outweigh the low risk of competitive
interactions over quality shelters (Newton and
Marquis 1982).
Echinometra lucunter is a nocturnal tropical
sea urchin found in west Atlantic coastal areas
from Florida to Brazil that abrades rock surfaces
with its teeth and spines to create protective
burrows (Hendler et al. 1995). Burrows offer
114
protection from predators and wave action, and
urchins rarely leave their burrows except to forage
on nearby algal patches at night (Shulman 1990).
E. lucunter often occur at high densities around
food patches ( 240 individuals m2, Metaxas and
Young 1998). Many habitats with abundant algal
resources and suitable substrate support highdensity urchin populations (Davidson and Grupa
2014), increasing competition for burrows.
Conspecific aggression over burrows is well
documented in E. lucunter (Grunbaum et al.
1978). Previous studies suggest that E. lucunter in
these environments of low burrow availability
commonly exhibit site fidelity, consistently
returning to the same burrow after a night of
feeding (McPherson 1969). Experiments have
shown that starved urchins are more likely to leave
burrows than those that have fed, and many
species forage only at night when predation risk is
low (Yanagisawa 1991). This suggests that leaving
burrows to forage requires risking predation and
that urchins may avoid leaving burrows unless the
benefit of feeding outweighs this risk.
E. lucunter occupy tidal pools formed in
limestone karst – an easily eroded limestone
landscape formed by dead coral – on the northern
shores of Little Cayman Island. These tidal pools
differ from sites of commonly studied urchin
Dartmouth Studies in Tropical Ecology 2014
aggregations in that they contain many empty
burrows of varying sizes. This high burrow
abundance may be caused by movement of urchins
to and from the ocean over time, leaving behind
unoccupied burrows that can be used by incoming
individuals. While site fidelity is commonly
observed in E. lucunter, this behavior could be a
product of the high urchin density and low burrow
availability where these urchins have been
traditionally studied. When the pressures of
burrow competition are reduced, E. lucunter
populations may display different burrow selection
behaviors. The high availability of suitable
burrows in Little Cayman tidal pools provides the
opportunity to examine these behaviors.
We investigated the burrow preferences and
between-burrow movements of E. lucunter where
burrow availability is high. We hypothesized that
urchins would not display site fidelity in
environments where suitable unoccupied burrows
are abundant. Urchins should also spend more
time in burrows that provide better protection from
predators and wave action. If deeper burrows
provide better protection, then urchin size and
burrow depth should not be related. Additionally,
urchins in deeper burrows should be less likely to
leave those burrows to forage and sacrifice their
high quality shelter. However, if burrows that fit
closely around an urchin provide more protection
by covering its body tightly, urchins should
occupy burrows of depths proportional to their
body size. In this case, urchins in burrows of
depths proportional to their body size should be
less likely to leave those burrows and risk
predation or wave damage.
METHODS
We examined patterns of E. lucunter burrow
occupation and movement in three limestone karst
tidal pools at Barge Dock, Little Cayman Island on
27-29 February 2014. On 27 and 28 February,
weather and surf conditions were mild. On 29
February, winds became stronger and wave
frequency and energy increased. We conducted all
movement observations under moderate tidal
conditions immediately preceding a new moon.
All three tidal pools were approximately 1 m2.
They all contained at least three times as many
unoccupied burrows as urchins and ranged in
burrow depths and urchin sizes. We defined
burrows as deep, cup-shaped indentations in the
limestone substrate of the tidal pools. Two of the
pools were approximately 2 m from the ocean, and
each was consistently replenished by waves
through an opening approximately 0.5 m wide.
Algal type and abundance, urchin density, urchin
size distribution, and burrow density were roughly
equivalent between these two pools. The third pool
was approximately 5 m from the ocean and was
infrequently replenished during our study period.
This pool had considerably lower algal cover,
urchin density, urchin size, and burrow density.
E. lucunter Burrow Occupation
To investigate the relationship between burrow
depth and E. lucunter body size, we surveyed 45
urchins in one pool 2 m from the ocean (nearocean) on 27 February (N = 26) and one pool 5 m
from the ocean (far-ocean) on 29 February (N =
19). We exhaustively surveyed urchins in half of
the near-ocean pool and all of the far-ocean pool.
We did not assume that urchins bored the burrows
in which they resided because it is very possible
that urchins actively abandon and exchange
burrows (Yanagisawa 1991). We removed urchins
from their burrows using forceps and measured
each urchin’s test diameter and full diameter
(including both test and spines, Fig. 1). We also
measured burrow depth, defined as the distance
from the flat rock surface surrounding the burrow
to the deepest point of the burrow (Fig. 2). After
we measured urchin size and burrow depth, we
returned urchins to their original burrows.
Figure 1. Body size measurements of sea urchin
E. lucunter.
115
Little Cayman
E. lucunter Movement
We examined E. lucunter movement and site
fidelity in the two pools each located 2 m from the
ocean on 27 and 28 February. On 27 February, we
marked 10 urchins from each pool (N=20) by
tagging a spine with a labeled piece of electrical
tape. For each tagged urchin, we measured the
depth of its burrow, placed a numbered marker at
the burrow, and took a photograph of the urchin
and its marker. We again measured urchin test
diameter and full diameter.
We returned to the two pools at 2230 and
noted whether each urchin had remained in its
original burrow or had moved elsewhere in the
tidal pool. On 28 February at 1530 we returned
and recorded which urchins had remained in their
original burrows and which had moved overnight.
RESULTS
Burrow occupation
Urchin body size was positively correlated with
burrow depth (Table 1, Fig. 3A, 3B). Urchin test
diameter ranged from 0.6 to 3.6 cm, and full
diameter ranged from 1.1 to 8.4 cm. Burrow depth
ranged from 0 to 7.7 cm.
Figure 2. Burrow depth was measured from lip of
burrow to its deepest point after an urchin was
removed.
A
Statistical Analyses and Modeling
To determine the burrow depth preferences of E.
lucunter, we performed a linear regression of
burrow depth by full urchin diameter in the two
pools sampled for burrow occupation. We then
used two-tailed t-tests to analyze patterns of E.
lucunter size (test diameter and full diameter) and
burrow depth between “movers” (urchins observed
outside their burrow at some point during the
observation period) and “non-movers” (urchins
never observed outside their original burrow).
We chose to measure the fit of an urchin to its
burrow depth by calculating the residuals of the
linear regression of burrow depth by full diameter.
For this analysis, we used only data from the nearocean pool in the burrow occupation, which was
more representative of the habitat type and size
distribution of urchins used for the movement
study. We assumed that urchins in appropriately
sized burrows would fall close to the regression
line, while urchins in burrows too large or small
for their bodies would show larger residuals. We
compared residuals between movers and nonmovers to evaluate whether burrow fit influenced
likelihood of urchin movement.
B
Figure 3. Burrow depth increased with test diameter (A) and full diameter (B).
116
Dartmouth Studies in Tropical Ecology 2014
move over the course of the night (Fig. 4, t = 3.37, P = 0.008, df = 9). Urchin body size (test
diameter or full diameter) was not related to
likelihood of movement from the original burrow
(Table 2). Non-mover burrow depth closely fit the
regression line of burrow depth to body size found
in the tide pool also surveyed for burrow
occupation (Fig. 5A).
Non-mover:
burrow depth = 0.60 *( full diameter) + 0.93
R2 = 0.70
Figure 4. Urchins in deeper burrows were less likely to
move than those in shallower burrows.
Movement
At 1530 on 28 February, we located 15 of the 20
urchins that we tagged at 1600 on February 27.
We did not locate 5 of the tagged urchins, possibly
because they left the pool, their tags fell off, or
they were depredated over the course of the night.
We identified seven of the 15 located urchins as
movers. Of these 7 movers, only one returned to
its original burrow the following day. The
remaining 8 urchins were non-movers. We
acknowledge the possibility that non-movers left
and returned to their burrows between observation
periods.
Urchins in deeper burrows were less likely to
A
(Equation 1)
Burrow occupation:
burrow depth = 0.55 *( full diameter) +1.18
R2 = 0.44
(Equation 2)
However, burrow depth and full diameter were not
related in mover urchins (Fig. 5B).
Mover:
burrow depth = 0.02 *( full diameter) + 2.65
R2 = 0.0005 (Equation 3)
Burrow occupation:
burrow depth = 0.55 *( full diameter) +1.18
R2 = 0.44
(Equation 2)
DISCUSSION
Our results suggest that in environments where
burrows are abundant, E. lucunter occupy burrows
B
Figure 5. (A) In non-movers, the positive correlation between full urchin diameter and burrow depth was
identical to that of all 26 urchins in the pool sampled (predicted correlation). (B) In movers, there was no
relationship between full urchin diameter and burrow depth. Solid line represents a linear regression of
urchins surveyed for burrow occupation in the refreshed pool. Points represent non-movers (A) or movers
(B). Dashed line represents a linear regression of burrow depth and full diameter in movers or non-movers.
117
Little Cayman
that fit tightly around them. Instead of using the
deepest burrows available, urchins occupied
burrows that were proportional to their body size
(Fig. 3). Urchins in burrows poorly fitted to their
body size were more likely to leave burrows to
forage (Fig. 5), suggesting that the shelter benefits
offered by these burrows did not outweigh the
potential benefits of feeding. It is also possible that
these urchins moved to actively seek out a better
burrow wave action or that wave action dislodged
them. Burrows with a tight fit around an urchin’s
body may be more protective, allowing the urchin
to dig its spines into the substrate (McClanahan
1988). Digging into a burrow could give an urchin
the ability to resist wave action and predation
more effectively than in a burrow that is either too
large or small. Alternatively, small urchins may
choose shallower burrows in order to avoid
invoking an aggressive response from larger
urchins. Deeper burrows may be more important
for larger urchins, and larger urchins are capable
of forcefully removing smaller urchins from these
burrows by initiating aggressive interactions
(Morishita 2009). Although this interaction may
be less likely in an environment where shelters are
abundant, it is possible that urchin burrow choice
is not plastic and does not change with shelter
availability in their environment.
The majority of movers (86%) did not return
to their original burrows. While previous studies
have reported site fidelity as common in E.
lucunter, these observations suggest that burrow
selection and fidelity may vary with environmental
conditions. Since these tidal pools have high
burrow availability, it may be less important for E.
lucunter to find and defend a single burrow.
Instead, they may leave a burrow at night to forage
and then occupy a similarly sized, closer burrow to
LITERATURE CITED
Grunbaum, H., G. Bergmann, D. Abbott, and J.
118
avoid the energy costs and predation risks of
returning to and defending their original burrow.
Finally, we found that non-movers occupied
deeper burrows and burrows that were more
closely proportional to their body size than movers
(Fig. 4, 5), but that body size did not differ
between movers and non-movers. These results
suggest that movers left burrows that were
proportionally too shallow for their bodies (and
therefore less protected from predators and waves)
in order to forage outside the burrow. Non-movers
remained in burrows that were proportional to
their body size, possibly because potential energy
gains from foraging do not outweigh the costs
associated with leaving a highly protective
burrow.
These results suggest that E. lucunter balance
burrow selection and foraging activities differently
depending on burrow availability. When available
burrows are abundant, urchins frequently move
among sites rather than displaying site fidelity, and
movement decisions appear to be influenced by
the protective capacity of a burrow. The variation
in how urchins choose burrows suggests that
studies of site selection, defense, and movement
may benefit from looking across a gradient of
resource limitation. This may help to better
understand the complex energetic trade-offs made
by organisms competing for limited resources.
ACKNOWLEDGEMENTS
We would like to thank CCMI for their hospitality
and Jessica Trout-Haney for her assistance with
data collection.
AUTHOR CONTRIBUTIONS
All authors contributed equally.
Ogden. 1978. Intraspecific agonistic behavior
in the rock boring sea urchin Echinometra
lucunter (L.). Bulletin of Marine Science 28:
Dartmouth Studies in Tropical Ecology 2014
181-188.
Hann, H.W. 1937. Life history of the Oven-bird in
southern Michigan. The Wilson Bulletin 49:
145-237.
Hendler, G., J.E. Miller, D.L. Pawson and P.M.
Kier. 1995. Sea stars, sea urchins, and allies:
echinoderms of Florida and the Caribbean.
Smithsonian Institution Press. Washington,
D.C.
Lewis, J.B. and G.S. Storey. 1984. Differences in
morphology and life history traits of the
echinoid Echinometra lucunter from different
habitats. Marine Ecology Progress Series 15:
207-211.
McClanahan, T. 1988. Coexistence in a sea urchin
guild and its implications to coral reef
diversity and degradation. Oecologia 77: 210218.
McPherson, B.F. 1969. Studies on the biology of
the tropical sea urchins Echinometra lucunter
and Echinometra viridis. Bulletin of Marine
Science 19:194-213.
Shulman, M.J. 1990. Aggression among sea
urchins on Caribbean coral reefs. Journal of
Experimental Marine Biology and Ecology
140: 197-207.
Switzer, P.V. 1993. Site fidelity in predictable and
unpredictable habitats. Evolutionary Ecology
7: 533-555.
Yanagisawa, T. 1991. Biology of Echinodermata.
CRC Press.
119
Little Cayman
A. PALMATA COLONY HEALTH AND SIZE AFFECT FISH ABUNDANCE, NOT SPECIES
DIVERSITY
ALEXA E. KRUPP, AMELIA L. RITGER AND ADAM N. SCHNEIDER
Faculty Editor: Celia Chen
Abstract: Reef building corals are integral to coral reef fish communities, providing fish with shelter and benthic
fauna upon which to feed. As global coral bleaching events destroy reefs and percent coral cover (health) decreases,
damaged benthic communities and degraded habitat structures may negatively impact fish abundance and diversity.
We examined how variation in live coral coverage and colony size of individual Acropora palmata colonies
influences resident fish assemblages at Bloody Bay Marine Park of Little Cayman Island. We surveyed and
compared fish communities at twenty-five A. palmata colonies of varying live coral coverage. We found that at the
colony level, coral health and colony size influences fish abundance, but not fish species diversity. Because species
did not have a preference for live coral coverage or colony size at the colony level, is important to consider scale
when making conservation decisions aimed at preserving species diversity. In coral reef systems, efforts to preserve
species diversity may require protection of the entire reef rather than individual colonies.
Key words: Acropora palmata, coral health, fish species diversity
INTRODUCTION
Reef building corals provide the structural
complexity and foundation of the benthic
community that numerous invertebrates and fish
rely on for shelter and food (Bell and Galzin
1984). The coral polyps form the foundation of the
benthic community that many fish rely on for food
while the calcium carbonate skeleton secreted by
corals contributes to the topography of reef
environments, providing fish with shelter and
hunting grounds (Lirman, 1999).
As the condition of reef-building corals is
negatively affected worldwide by anthropogenic
inputs and environmental stressors induced by
climate change, the health of the entire reef
community is also harmed. Fish abundances have
been disproportionately negatively impacted by
loss of live coral coverage and are sensitive to as
little as 5% loss of coral coverage across a reef
(Feary et al. 2007, Bell and Galzin 1984). Fish
diversity is also highly vulnerable to coral loss. Up
to 80% of fish species may be lost in reefs with
coral die-offs (Pratchett et al. 2011).
Fish vary in their reliance on coral colonies as
a function of their functional feeding group.
Herbivores might use corals as shelter, while
invertivores and carnivores might use corals as
hunting grounds. Such differences suggest that
although entire fish communities are affected by
changes in coral colony health, these impacts vary
by functional group (Pratchett et al. 2011).
120
Acropora palmata (elkhorn coral) is one of the
dominant species comprising overall coral
coverage and contributing to the topography of
shallow reef environments (Lirman 1999). We
examined how fish assemblages residing in
elkhorn coral changed with varying degrees of
host colony health in a Caribbean fringing reef.
First, we tested the hypothesis that fish abundance
and diversity are affected by the size and percent
live coverage of elkhorn coral. If live coverage
alone predicts fish communities, fish may be
primarily attracted to the benthic fauna on the
coral. If fish abundance and diversity are solely a
function of colony surface area, fish may inhabit
the coral primarily for shelter. Finally, if fish are
dependent on both the benthic community and
habitat structure provided by elkhorn coral, we
would expect both size and percent live coral
coverage to positively influence fish abundance
and diversity.
We also examined the effect of live coral
coverage on fish abundance and diversity by
functional group. Since functional feeding groups
have been found to respond differently to
disturbance (Pratchett et al. 2011), we predicted
that there would be variation among these groups
in response to changes in live coral coverage.
Differences in fish community composition with
live coverage would indicate that particular
functional groups or species may be particularly
susceptible to coral loss.
Dartmouth Studies in Tropical Ecology 2014
METHODS
We sampled along the fore reef of Bloody Bay
Marine Park in Little Cayman Island on 27
February from 9:00 to 17:30. We selected twentyfive A. palmata colonies along a transect parallel
to shore, excluding colonies less than 1m long and
those with extensive macroalgal, octocoral or
hydrozoan coverage. We measured the height and
length of each colony and calculated twodimensional colony area. We estimated live tissue
coverage within six percentage classes (dead, 120%, 21-40%, 41-60%, 61-80%, 81%-100%).
We watched each colony of A. palmata for
two minutes, during which time we recorded fish
abundance and identified species of fish located
within ½ meter away of the focal colony and
present for the full two minutes. We categorized
observed fishes by functional feeding groups as
herbivores, invertivores, or carnivores.
Statistical analysis
We calculated fish species diversity at each colony
using Simpson's Diversity Index. We used
Analysis of Variance (ANOVA) to examine
whether fish abundance or species diversity
changed with live coral coverage class. We
performed an ANOVA to determine if coral
colony size had an effect on live coral coverage.
We performed an Analysis of Covariance
(ANCOVA) of fish abundance and diversity with
respect to coral coverage and coral size. Finally,
we used an ANOVA to examine the relationship
between reef fish abundance by functional feeding
group and live coral coverage. All analyses were
performed using Al Young Biodiversity Calculator
and JMP® 10 statistical software (SAS Institute,
Cary, NC).
Figure 1. Fish abundance increased as area
increased on coral colonies containing some level of
live coral coverage. Larger dead corals had less fish
than smaller dead corals. No regression line is
presented for size class 4 due to low sample size.
RESULTS
We observed a total of twenty fish species from
ten families on A. palmata colonies (Appendix 1).
On average, we found six resident fish individuals
from four different species at each colony. Live
coral coverage class affected fish abundance
(F4,24= 5.47, P = 0.0073, Fig. 1) but did not affect
species diversity (F5, 22 = 0.23, P = 0.942; Fig. 2).
A. palmata colony size ranged from 0.48 m2 to
3.38 m2 and showed no relationship with live coral
coverage (F5, 42 = 1.34, P = 0.289). Although coral
colony size did not affect fish diversity (slope = -
Figure 2. Reef fish species diversity did not change with A. palmata percent live coral cover.
121
Little Cayman
Figure 3. Reef fish functional feeding group was not associated with A. palmata percent live coral coverage.
Herbivores and invertivores were most abundant in dead A. palmata and carnivores were never found in dead A.
palmata.
0.047 ± 0.074, P = 0.53, r2 = 0.019), coral colony
size was a predictor of fish abundance (slope =
2.13 ± 1.1, P = 0.058, r2 = 0.15). The interaction of
live coral coverage and coral size was a significant
predictor of fish abundance (F4, 42 = 0.002, P =
0.02; Fig. 1) but not fish diversity (F4, 42 = 1.16, P
= 0.37). Fish abundance by functional feeding
group did not change with live coral coverage
(Carnivores: F5, 24 = 0.41, P = 0.84; Invertivores:
F5, 24 = 1.20, P = 0.35; Herbivores: F5, 24 = 2.68, P
= 0.053; Fig. 3).
DISCUSSION
The health of coral reefs may have substantial
effects on resident fish communities. We found
that health and size individual A. palmata coral
colonies influenced fish abundance but not species
diversity. Greater numbers of fish were found
around healthier, larger corals, but species
composition did not change across coral colonies
(Fig. 1). This suggests that fish communities are
dependent upon corals that provide both habitat
structure and a benthic community.
We found no meaningful difference between fish
abundance within each functional feeding group
across live coral coverage. There were, however,
slightly higher herbivore abundances on dead
corals (Fig. 3), which could be the result of
increased grazing opportunities from higher algal
coverage on dead coral.
Although more fish congregated around
larger, healthier coral colonies, many different fish
species from all functional feeding groups could
122
be found at individual live colonies of varying
health and sizes (Fig. 3). It may be that A. palmata
live coral coverage influenced fish abundance but
not fish diversity because changes in diversity
cannot be seen by examining a single coral colony;
rather, impacts of coral health and coral size on
fish diversity must be examined over an entire
reef. When the overall health of a reef is
significantly compromised, reef fish diversity is
negatively impacted as dead reefs lose both habitat
structure and benthic communities (Alvarez-Filip
et al 2009).
These conclusions have important
implications for our understanding of coral reef
ecology and conservation. Fish species did not
prefer any particular level of live coral coverage
class, indicating that health of individual colonies
on a single reef may not impact the distribution of
fish species as much as coral health of an entire
reef. Therefore, reef preservation efforts aimed at
promoting healthy and diverse fish communities
should focus on preserving the presence of corals
on a reef instead of individual colony health.
ACKNOWLEDGEMENTS
We would like to thank Celia Chen, Vivek
Venkataraman, and Jessica Trout-Haney for their
valuable feedback on our paper. We would also
like to thank the damselfish for keeping us in line
as we invaded their territory.
AUTHOR CONTRIBUTIONS
All authors contributed equally.
Dartmouth Studies in Tropical Ecology 2014
LITERATURE CITED
Alverez-Filip, L., Dulvy, N.K, Gill, J.A., Cote,
I.M., Watkinson, A.R. 2009. Flattening of
Caribbean coral reefs: region-wide declines in
architectural complexity. Proceedings of the
Royal Society 276: 3019-3025.
Bell, J.D., and Galzin, R. 1984. Influence of live
coral cover on coral-reef fish communities.
Marine Ecology Progress Series 15: 265-274.
Feary, D., Almany, G., McCormick, M., and
Jones, G. 2007. Habitat choice, recruitment
and the response of coral reef fishes to coral
degradation. Oecologia 153: 727-737.
Graham, N. A. J., et al. "Coral mortality versus
structural collapse as drivers of corallivorous
butterflyfish decline." Biodiversity and
conservation 18.12 (2009): 3325-3336.
McCormick, M.I. 1995. Fish feeding on mobile
benthic invertebrates: influence onf spatial
variability in habitat associations. Marine
Biology. 121: 627-637.
Patterson, K. L., J. W. Porter, K. E. Ritchie, S. W.
Polson, E. Mueller, E. C. Peters, D. L.
Santavy, and G. W. Smiths. 2002. The
etiology of white pox, a lethal disease of the
Caribbean elkhorn coral, Acropora palmata.
Proceedings of the National Academy of
Sciences of the United States of America 99:
8725-8730.
Pratchett, Morgan S, A. S. Hoey, S. K. Wilson, V.
Messmer, and N. A. Graham. 2011. Changes
in biodiversity and functioning of reef fish
assemblages following coral bleaching and
coral loss. Diversity 3.3: 424-452.
123
Little Cayman
THE ROLE OF HETEROTYPIC SCHOOLING IN FORAGING BEHAVIOR OF PARROTFISH
EMILY R. GOODWIN AND AARON J. MONDSHINE
Faculty Editor: Celia Y. Chen
Abstract: Mixed-species aggregations offer organisms the benefits of anti-predatory defense while reducing
competition between species through reduced niche overlap. Coral reef fish are often found foraging or resting in
mixed-species aggregations, or heterotypic schools, near reef structures. We explored two aspects of heterotypic
schools in reef fish: first, whether heterotypic school size varies with school behavior, and second, how participation
in heterotypic schools affects foraging rate in parrotfish. We found that school behavior impacted school size but not
species richness. Resting schools had significantly higher fish abundance than foraging schools. In addition,
parrotfish had increased foraging rates within heterotypic aggregations than when foraging alone. However, for
parrotfish foraging in aggregations, foraging rates decreased with increased group size. This suggests that
moderately sized foraging groups are optimal for parrotfish. Heterotypic schooling, in contrast to homotypic
schooling, allows parrotfish to benefit from increased safety from predators without sacrificing foraging rate due to
competition. The tradeoff between protection from predation and feeding competition could play a role in the
evolution of mixed-species associations.
Key words: heterotypic schooling, mixed-species aggregations, parrotfish, schooling behavior
INTRODUCTION
In many organisms, the advantages of social
groups include benefits related to reproduction,
foraging, and predator avoidance. While many of
these organisms are found in conspecific groups,
mixed-species aggregations have also been
observed in a variety of birds, primates, and fish
(Morse 1977). These aggregations are also thought
to confer advantages related to foraging or
predator evasion, yet differ from single-species
aggregations in that feeding competition is
reduced in mixed-species aggregations due to
reduced niche overlap between species (Ehrlich
and Ehrlich 1973, Morse 1977, Lukoschek and
McCormick 2000).
Teleost fish schools in reef habitats represent a
classic example of both single- and mixed-species
aggregations (Pitcher and Parrish 1993). Fish
occurring in homotypic, or single-species, schools
may receive the anti-predatory benefits of
increased group size. Heterotypic, or mixedspecies, schools allow individual fish to decrease
predation risk while minimizing costs of feeding
competition through mechanisms such as niche
partitioning. However, school behavior may also
impact this trade-off between predator avoidance
and competition. Coral reef fish often aggregate as
heterotypic schools that rest or forage near
underwater structures (Itzkowitz 1977, Ogden and
Ehrlich 1977). Since there are no foraging benefits
124
for a resting aggregation, these groups are mainly
driven by the need to maximize group size as an
anti-predator defense. In contrast, foraging
aggregations offer the benefits of group-driven
anti-predator defense at the cost of increased
feeding competition between species. This creates
a situation in which resting schools may benefit
from larger group size regardless of species,
whereas foraging schools are driven to decrease
competition through smaller groups with higher
species diversity and decreased niche overlap.
This study assesses whether these patterns are
occurring in heterotypic schools of Caribbean reef
fish on Little Cayman Island. We hypothesize that
resting schools will have higher overall fish
abundance, whereas foraging schools will have
smaller schools with increased species diversity.
As a case study, we also examine how heterotypic
schooling benefits parrotfish (Scaridae) on Little
Cayman Island. Parrotfish are a common
Caribbean reef fish that participates in mixedspecies aggregations. Their feeding strategy
requires rapid and noisy biting that can expose
individuals to predation and incite aggression by
notifying potential competitors of a food source
(Ogden and Buckman 1973, Overholtzer and
Motta 2000). Parrotfish can minimize these risks
by schooling with conspecifics, but this increases
feeding competition (Overholtzer and Motta
2000). Participation in heterotypic schools may
Dartmouth Studies in Tropical Ecology 2014
METHODS
Data Collection
We observed 22 heterotypic aggregations and 53
parrotfish (32 in aggregations; 21 solitary) at
Pirates Point, Cumber’s Cove, and Grape Tree
Bay, Little Cayman, on 5-7 March 2014. We
opportunistically observed aggregations at these
sites and measured total fish abundance, species
richness, species abundances, and school behavior.
School behavior was determined over a 5-minute
period and was characterized as either foraging or
resting. Schooling behavior was found to be
consistent over the entire 5-minute period. We also
opportunistically measured foraging rates (bites in
1 minute intervals) of both solitary parrotfish and
those in heterotypic schools.
Statistical Analysis
We used a principal components analysis (PCA) to
compare the species composition of aggregations.
We used a one-tailed t-test to test whether species
richness and fish abundance differed by school
behavior. We also used a t-test to compare the
foraging rates of solitary parrotfish and those in
heterotypic schools. For parrotfish in schools, we
used a linear regression to test whether school size
affected the foraging rates of individual parrotfish.
Fish abundance
200
150
100
50
0
Foraging
Resting
Figure 1. Fish abundance was significantly
higher within resting aggregations than within
foraging aggregations.
RESULTS
We observed 20 different species of fish, seven of
which were species of parrotfish. Haemulon
carbonarium was the most abundant species,
followed by Acanthurus coeruleus, Kyphosus
sectatrix, and Haemulon flavolineatum (Fig. A).
At least three parrotfish individuals were found in
all of the schools. We did not observe any
homotypic schools of parrotfish.
7
Species richness
therefore confer protection from predation while
minimizing competition with other parrotfish.
Therefore, we predict that parrotfish in heterotypic
schools will have higher feeding rates than solitary
parrotfish, since they can afford to invest less
energy in predator detection and more energy
foraging.
6
5
4
3
2
1
0
Foraging
Resting
Figure 2. There was no significant difference
in species richness between foraging and
resting aggregations.
The PCA showed no clear trend with respect to
composition of fish aggregations. Principal
components 1 and 2 accounted for 16.1% and
15.6% of the total variance, respectively. The
loading patterns of the eigenvectors did not appear
to have any biological significance.
We found that fish abundance was
significantly higher in resting aggregations (mean
± SE = 123.90 ± 21.05) than in foraging
aggregations (mean ± SE = 50.40 ± 21.047, t1 =
2.47, P = 0.012; Fig. 1). However, there was no
difference observed for species richness between
resting (mean ± SE = 4.40 ± 0.60) and foraging
aggregations (mean ± SE = 5.50 ± 0.60, t1 = 1.30,
P = 0.11, Fig. 2).
Foraging rate was significantly higher for
parrotfish in aggregations (mean ± SE = 20.69 ±
1.51 bites/min) than for solitary parrotfish (mean ±
SE = 11.81 ± 1.87 bites/min, t1 = 3.69, P = 0.0003;
Fig. 3). Parrotfish in mixed-species aggregations
foraged less efficiently with increasing
aggregation size (slope= -0.050 ± 0.022, P =
0.033, r2 = 0.14; Fig. 4).
125
25
50
20
40
Foraging Rate
(bites/minute)
Foraging rate
(bites/minute)
Little Cayman
15
10
5
0
20
10
0
0
Within Aggregation
Solitary
Figure 3. Foraging rate for parrotfish within an
aggregation was significantly higher than for solitary
parrotfish.
DISCUSSION
Heterotypic aggregations may provide increased
protection from predators while minimizing the
costs of conspecific competition. However, this
tradeoff may be influenced by school behavior,
species composition, and aggregation size. We
found that resting schools had higher overall fish
abundance than foraging schools (Fig. 1). This
could be a result of (1) the anti-predatory benefits
of resting in a large group, or (2) the habitat
structure of a location being particularly suitable
for a wide range of fish species. Because there is
no feeding competition in resting schools, there is
no tradeoff between protection from predation and
feeding competition. We observed no difference in
species richness between resting and foraging
schools (Fig. 2), indicating that the pressures of
conspecific competition are not necessarily driving
greater species diversity in foraging schools
compared to resting schools.
Parrotfish exhibited behavior consistent with a
tradeoff between predation risk and feeding
competition. Foraging rate was significantly
higher for parrotfish in aggregations as compared
to those feeding alone (Fig. 3). This could be a
result of either increased predator detection in a
group setting, which allows individuals to devote
more time to foraging, or group facilitation in
locating patchily distributed resources. Increased
predator detection is more likely because the
macroalgae parrotfish foraged on was widespread.
However, for parrotfish in aggregations, foraging
LITERATURE CITED
Ehrlich, P.R., and A.H. Ehrlich. 1973.
126
30
100
200
300
Aggregation size
Figure 4. Foraging rate (bites/minute) decreased
significantly with increased aggregation size.
rate decreased as school size increased. This could
be explained
by increased feeding competition in
\
larger groups. Therefore, parrotfish in moderately
sized groups have the highest foraging rates. Since
parrotfish in small aggregations have higher
feeding rates than both parrotfish in large
aggregations and solitary parrotfish, joining small
aggregations appears to maximize foraging
success.
Our results suggest that the formation of
heterotypic schools may be driven by both
foraging benefits and predation risk. Parrotfish
were never found in homotypic schools, but were
found in all observed heterotypic schools. This
suggests that parrotfish in homotypic schools may
have an even smaller optimal group size than those
in heterotypic schools due to increased feeding
competition. Heterotypic schooling allows
parrotfish to benefit from increased safety without
sacrificing their foraging rate due to competition.
This anti-predatory-competition tradeoff could
play a role in the evolution of mixed-species
associations.
ACKNOWLEDGEMENTS
We are indebted to C. Chen, V. Venkataraman,
and J. Trout-Haney for their assistance at all stages
of this study. We would also like to thank the FSP
group for their valuable feedback.
AUTHOR CONTRIBUTIONS
Both authors contributed equally in the data
collection and writing process.
Coevolution: heterotypic schooling in
Caribbean reef fishes. The American
Dartmouth Studies in Tropical Ecology 2014
Naturalist 107: 158-160.
Foster, S.A. 1985. Group foraging by a coral reef
fish: a mechanism for gaining access to
defended resources. Animal Behavior 33: 782792.
Itzkowitz, M.1977. Social dynamics of mixedspecies groups of Jamaican reef fishes.
Behavioral Ecology and Sociobiology. 2: 361384.
Lukoschek, V., and M.I. McCormick. 2000. A
review of multi-species foraging associations
in fishes and their ecological significance.
Proceedings of the International Coral Reef
Symposium. Bali, Indonesia 23-27.
Morse, D.H. 1977. Feeding behavior and predator
avoidance in heterospecific groups.
BioScience 27: 332-339.
Ogden, J.C., and N.S. Buckman. 1973.
Movements, foraging groups, and diurnal
migrations of the striped parrotfish Scarus
croicensis. Ecology 54: 589-596.
Ogden, J.C., and P.R. Ehlrich. 1977. The behavior
of heterotypic resting schools of juvenile
grunts (Pomadasyidae). Marine Biology 273280.
Overholtzer, K.L., and P.J. Motta. 2000. Effects of
mixed-species foraging groups on the feeding
and aggression of juvenile parrotfishes.
Environmental Biology of Fishes 58: 345-354.
Pitcher, T.J., and J.K. Parrish. 1993. Functions of
shoaling behavior in teleosts. In: Pitcher T.J.
(ed.), Behavior of Teleost Fishes. Chapman
and Hall, London: 363-439.
127
Little Cayman
SALT POND MINNOWS ACCLIMATIZE TO SALINITY
REBECCA C. NOVELLO, AMELIA L. RITGER, AND ZACHARY T. WOOD
Faculty editor: Celia Chen
Abstract: Acclimatization allows organisms to adjust to changes in their environment. Many fish can adjust to
changes in components of water chemistry such as salinity. Ponds on Little Cayman Island are subject to large
salinity ranges due to evaporation, seasonal rainfall, and marine water influxes. We examined how Gambusia spp.
minnows are affected by variations in salinity levels and whether minnows have acclimatized to the salinity of their
native ponds. We exposed fish from two ponds (high and low salinity) to a gradient of salinity treatments and
calculated the 24-hour 50% lethal concentration (LC50). We performed a laboratory-based reciprocal transplant
between fish of the two ponds and measured stress (via percent activity) to assess the effects of acclimatization to
native-pond salinity upon salinity tolerance. Minnows from the high salinity pond had a higher upper salinity
tolerance than those from the low salinity pond. Minnows were more stressed in water from the foreign pond than
from their native pond. We conclude that minnows acclimatize to the salinity of their native ponds. Salinity levels
higher or lower than that of native ponds results in impaired performance.
Key words: acclimatization, Gambusia, LC50, minnows, salinity, tolerance
INTRODUCTION
Organisms that can physiologically adjust to
changes in their environment by acclimatization
can be more resilient to environmental change
(Somero 2009). In aquatic ecosystems, organisms
may acclimatize to changes in temperature,
oxygen, pH, and salinity. Salinity, in particular,
influences development, growth, and metabolism
of aquatic organisms such as fish (Boeuf and
Payan 2001). Fish use active transport, specialized
chloride cells, and urination patterns to tolerate
changes in salinity (Evans 2002, Uchida et al
2000, Loretz and Bern 1980). Fish can show
varying levels of tolerance to changes in water
chemistry; for example, estuary systems contain
euryhaline killifish that are highly tolerant to
changes in water salinity and therefore have a
broader spatial range in the estuary than
stenohaline species (Griffith 1974).
Shallow inland ponds near seacoasts have
large seasonal fluctuations in salinity due to
evaporation, seasonal rainfall, and marine water
influxes, and variation in basin morphometry may
result in differences in salinity across ponds. Little
Cayman Island contains many such ponds with
differing salinities. During the dry season (January
- April), local fishermen have reported salinities
reaching 70 ppt (Tom Provost, personal
communication). Despite variation in salinity
among ponds, many contain populations of
Gambusia spp. minnows. Therefore, these
128
minnows must have some mechanism for dealing
with variation in salinity.
We studied how fluctuations in native pond
salinity affect salinity tolerance in Gambusia
minnows. We tested whether minnows from a high
salinity environment had higher upper salinity
tolerances than minnows from a low salinity
environment. If minnows wereacclimatized to the
salinity of their environment, minnows from a
high salinity environment would be more stressed
than minnows from a low salinity environment
when exposed to low salinities. Similarly,
minnows from a low salinity environment would
be more stressed than minnows from a high
salinity environment when exposed to high
salinities.
METHODS
We collected 150 minnows and 80 litres of water
from two ponds of approximately one meter depth
on Little Cayman Island on 6 March, 2014.
Jackson's Pond, located on the north side of the
island, had a surface area of about 110,000 m2 and
a salinity of 35 ppt at the time of the study. Tarpon
Lake, located on the south side of the island, had a
surface area of about 190,000 m2 and a salinity of
52 ppt at the time of the study.
Salinity tolerance
We experimentally assessed the upper salinity
tolerances of fish from Jackson's Pond and Tarpon
Dartmouth Studies in Tropical Ecology 2014
Lake. At 16:30 on 6 March, we placed 10
minnows of similar size from each pond into one
of 10 salinity treatments ranging from 52 ppt to 70
ppt at 2 ppt intervals. We altered salinity of
treatments by adding NaCl to 3 L of water from
the native pond and assessing salinity with a
refractometer. We also placed 10 minnows from
each pond into a control treatment of 3 L of water
from the native pond without salt additions.
We recorded minnow mortality at 40-minute
intervals from 16:30 to 23:30 on 6 March and
04:30 to 16:30 on 7 March. For minnows from
each pond, we fit a logistic model to proportional
mortality (M) across salinities at 24 hrs.
(Equation 1)
This model has an upper asymptote at one and a
lower asymptote at zero. As ß1 becomes more
negative, the slope of the curve increases. The
model is centered around point ß2, which is when
M = 0.5 and is therefore the 50% lethal
concentration (LC50) for salt. ß3 is the difference
in LC50s between the high salinity (Tarpon) lake
and low salinity (Jackson) pond. We compared
LC50s between minnows from the two pond
sources to determine whether upper salinity
tolerances differed between ponds. We performed
all analyses in R.
Reciprocal transplant
We conducted a reciprocal transplant experiment
to determine the acclimatization of minnows to the
salinity of their native ponds. We placed twelve
minnows from each pond into 6 L of water from
their own pond (native) and twelve into 6 L of
water from the other study pond (foreign). We
gave the minnows two hours to acclimate to
treatments. After the treatment acclimation period,
we moved two minnows from haphazardly chosen
treatments into 150 mL of the same treatment
water every fifteen minutes and allowed them to
acclimated for another fifteen minutes. We
measured percent activity for the two minnows in
those treatments. We defined percent activity as
the percent of time spent swimming over the
course of a minute and used it as a proxy for
minnow stress, with more movement indicating
higher stress. We compared percent activity across
water and minnow sources using a two-way
ANOVA. We performed this analysis in R.
RESULTS
Salinity tolerance
Tarpon Lake minnows had a greater upper salinity
tolerance (LC50) than Jackson's Pond minnows
(Fig. 1). LC50 for Jackson's Pond minnows was
60.0 ppt, while that of Tarpon Lake minnows was
71.2 ppt (Table 1, P < 0.0001). 83% of all minnow
deaths occurred in the first 4 hours of the study.
Figure 1. Minnows from a low salinity pond (Jackson’s Pond, 35 ppt) died at lower salinities than fish
from a high salinity pond (Tarpon Lake, 52 ppt).
129
Little Cayman
Table 1. Regression coefficients for logistic model
Coefficient
Est. ± Std. Err.
P
β1
-0.44 ± 0.10
0.00038
β2
60.0 ± 0.6
< 0.0001
β3
11.2 ± 1.2
< 0.0001
Reciprocal Transplant
Minnows were more active in water from the
foreign pond than their native pond (Figure 2,
Table 2, P = 0.009). Neither group of fish was
more active, and neither water source caused
greater fish activity (Table 2).
Figure 2. Minnows were more active in water from
the opposite pond than from their own pond.
Table 2. ANOVA for fish activity on water and fish
source
Mean
F
Effect
df
Squares
P
Water
1
0.085
0.734
0.40
Fish
1
0.002
0.017
0.90
Water *
Fish
1
0.863
7.451
0.0091
Residuals
44
0.116
130
DISCUSSION
We tested whether minnows acclimatize to
different salinities by examining two types of
responses: changes in upper lethal salinity levels
and acclimatization to a particular salinity.
Minnows in more saline lakes had higher LC50s,
indicating that they have a higher upper salinity
tolerance (Fig. 1). For both of our study ponds,
minnows could tolerate salinities higher than those
in their native ponds (Fig 2). This demonstrates
that as salinity increases, minnows develop an
increased salinity tolerance that buffers them
against further rapid increases in salinity.
However, as salinity increases, this buffer zone
between experienced salinity and tolerable salinity
decreases (i.e., from 25 ppt for Jackson’s Pond to
19 ppt for Tarpon Lake), suggesting that fish
cannot increase their salinity tolerance ad
infinitum.
Fish were more active when placed in water of
different salinity than that of their native pond
(Fig. 2). Acclimatization to more saline water
appears to cause fish to show poorer performance
in low salinity water, and vice versa. High activity
is metabolically costly. Therefore, fish that move
will have decreased energy reserves, which could
lead to decreased long-term fitness. Although
minnows may not die within 24 hours at these
lower or higher salinities, they may still exhibit
stress responses.
The ability to acclimatize to changes in
salinity is important in light of anthopogenic
effects. Lakes in cold temperate regions are
becoming more saline due to road salt influxes
(Novotny et al. 2008). Humans have altered the
flow regimes in estuaries, changing where the
boundaries between marine and freshwater occur
(Colonnello and Medina 1998). Furthermore, wet
and dry seasons are expected to become more
pronounced (IPCC 2007), which will affect the
degree to which salinity fluctuates seasonally in
inland saltwater lakes and ponds. Therefore,
selection for organisms that can acclimatize to
changing salinity should become stronger in many
different aquatic systems. Although many
organisms can increase their salinity tolerance, our
study suggests that this increase probably has
limits. Furthermore, we have shown that the
acclimatization of organisms to one salinity level
can decrease functionality at other salinities,
suggesting that acclimatization in areas of
Dartmouth Studies in Tropical Ecology 2014
increasing salinity may limit the ability of
organisms to leave these areas.
ACKNOWLEDGEMENTS
We would like to thank the Central Caribbean
Marine Institute and their Little Cayman Research
Centre for their hospitality and lab resources. We
LITERATURE CITED
Boeuf, G., and Payan, P. 2001. How should
salinity affect fish growth? Comparative
Biochemistry and Physiology Part C:
Toxicology and Pharmacology 130: 411-423.
Colonello, G. and Medina, E. 1998. Vegetation
changes induced by dam construction in a
tropical estuary: the case of the Manamo
River, Orinoco Delta (Venezuela). Plant
Ecology. 139.2: 145-154.
Griffith, R.W. 1974. Environment and salinity
tolerance in the genus Fundulus. Copeia 1974:
319-331.
Evans, D.H. 2002. Cell signaling and ion transport
across the fish gill epithelium. Journal of
Experimental Zoology 293: 336-347.
International Panel on Climate Change. 2007.
Climate Change 2007: Synthesis Report.
Loretz, C.A., and Bern, H.A. 1980. Ion transport
by the urinary bladder of the gobiid teleost,
would also like to thank Celia Chen, Jessica TroutHaney and Vivek Venkataraman for driving us
around Little Cayman and editing our manuscript.
AUTHOR CONTRIBUTIONS
All authors contributed equally.
Gillichthys mirabilis. American Journal of
Physiology- Regulatory, Integrative, and
Comparative Physiology 239: 415-423.
Novotny, E.V., Murphy, D., and Stefan, H.G.
2008. Increase of urban lake salinity by road
deicing salt. Science of the Total Environment
406: 131-144.
Somero, G.N. 2009. The physiology of climate
change: how potentials for acclimatization and
genetic adaptation will determine ‘winners’
and ‘losers.’ Journal of Experimental Biology
213: 912-920.
Uchida, K., Kaneko, T., Miyazaki, H. Hasegawa,
S., and Hirano, T. 2000. Excellent salinity
tolerance of Mozambique tilapia
(Oreochromis mossambicus): elevated
chloride cell activity in the branchial and
opercular epithelia of the fish adapted to
concentrated seawater. Zoological Science 17:
149-1600.
131
Little Cayman
ENERGY ALLOCATION VARIES THROUGHOUT GROWTH AND BETWEEN SEXES OF
PTEROIS VOLITANS
MAYA H. DEGROOTE AND ALEXA E. KRUPP
Faculty Editor: Celia Chen
Abstract: The rapid invasion of lionfish (Pterois volitans) in the Caribbean poses significant threats to fish
communities through competition with native piscivores. Understanding the life history strategies of lionfish, which
include rapid growth rate and high fecundity, may help explain the mechanisms underlying their competitive
advantage. In Little Cayman, we examined how spine, fin and body length changed with body weight and how
investments in fat and gonad mass differed between mature males and females. As lionfish weight increased, their
relative spine, fin, and body length decreased. Both female and male lionfish invested more in relative gonad weight
as their body weight increased, but only males exhibited increasing fat content with body size. These findings
suggest that as lionfish grow, their reproductive capability continues to increase and larger males with more energy
reserves may have a greater ability to withstand stressors. As increasingly larger lionfish are found throughout their
non-native ranges, higher reproductive and survival capabilities in larger adults may contribute to a higher rate of
lionfish proliferation.
Key words: invasive, lionfish, Pterois volitans, resource allocation
INTRODUCTION
Lionfish (Pterois volitans), native to the IndoPacific were introduced to southern Florida in the
early 1990s and have spread prolifically
throughout the Caribbean (Albin & Hixon 2008).
As top predators, lionfish can adversely affect reef
ecosystems through competition with native
piscivores and consumption of a variety of native
reef fishes, crabs and shrimps (Hare & Whitfield
2003). The competitive advantage of this species
may be attributed to their rapid growth rate, high
fecundity, and venomous spines (Morris et al.
2011).
All organisms face trade-offs in how resources
are allocated toward reproduction, defense, and
growth. The principle of allocation (Cody 1966)
suggests that organisms evolve to allocate
resources in ways that maximize their fitness.
Juvenile and adult fish allocate resources
differently based on the predatory threats and
reproductive demands associated with their life
stages. Juvenile fish prioritize rapid growth in part
to escape predation, whereas adults invest more
heavily in traits relating to reproductive success.
Juvenile lionfish may invest more in
morphological features such as long venomous
spines and protruding fan-like pectoral fins that
increase their apparent size and likely reduce
predation. Female and male lionfish may also have
different strategies of resource allocation due to
132
differing reproductive energy requirements.
Mature female lionfish in the Caribbean can
reproduce multiple times per month (Morris et al
2011); this short, frequent reproductive cycle is
energy intensive, and may explain why females
have less body fat in relation to total body length
than males (Costello et al. 2012).
We studied how lionfish invest in defensive
characteristics, fat storage, and reproduction
throughout ontogeny and between sexes. Since
lionfish exhibit indeterminate growth, allocation to
morphological features may vary with size,
particularly before and after sexual maturity. We
tested the hypothesis that smaller and more
vulnerable juvenile lionfish invest more heavily in
morphological features related to defense relative
to larger lionfish. Alternatively, these
morphological traits such as spine length and
pectoral fin length grow isometrically with body
size. Due to the frequency in which females
reproduce we also predicted that females allocate
more resources to reproduction than males and
will thus have lower relative body fat and larger
gonad size.
METHODS
We measured twenty-seven juvenile lionfish
collected in February-March 2014 from Mary’s
Bay, Pirate’s Point Dock, and South Hole Sound
in Little Cayman. We also dissected ten lionfish
Dartmouth Studies in Tropical Ecology 2014
culled from Joy’s Bay on March 5, 2014. We
weighed all lionfish and measured total body
length, pectoral fin length, and both dorsal and
pectoral spine length. Total length was measured
as the distance from the mouth to the tip of the
caudal fin. Pectoral fin length was measured from
the base of the fin to its longest point. The longest
dorsal spine and the first pectoral fin were
measured using SPI 2000 calipers.
In mature fish, we determined sex and then
removed and weighed gonads. We also determined
body fat mass by removing and weighing all fat in
the body cavity. All weights were made using an
AdventurerTM OHAUS. We also used data
collected in Little Cayman in March 2012 by
Costello et al. (2012) in our analysis.
Statistical analysis
All morphological measurements and the total
body weight of lionfish were log transformed. We
used regressions to examine the relationship
between morphological traits and body weight. If
the length of a feature plotted against weight in
log-log space exhibited a slope >1 (positive
allometry), this indicates that the feature increases
disproportionately relative to body size, whereas a
slope < 1 (negative allometry) indicates that the
feature decreased disproportionately relative to
body size. If the length of a feature was
proportional to body weight, the slope = 1
(isometric relationship).
To examine resource allocation throughout
ontogeny, we performed a regression analysis of
body length, pectoral fin length, pectoral spine
length, and dorsal spine length, with respect to
body weight. To test for differences between sexes
in allocation toward body fat and reproduction, we
regressed fat and gonad mass with weight in males
and females. We performed an analysis of
Figure 1. Lionfish length had a negative allometric
(slope = 0.28 ± 0.02) relationship with weight.
covariance (ANCOVA) of gonad mass and fat
mass in males and females. All analyses were
performed using JMP® 10 statistical software
(SAS Institute, Cary, NC).
RESULTS
The body weight of lionfish ranged from 0.6 g to
720.1 g and body length ranged from 57 mm to
381 mm. Lionfish total body length scaled in a
negatively allometric fashion with weight,
indicating the rate of body length increase was
lower than that of weight gain (slope = 0.28 ±
0.02, P < 0.0001, r2 = 0.99; Fig. 1). Likewise,
length of the pectoral fin, dorsal spine and pectoral
spine scaled negatively with lionfish weight (slope
= 0.20 ± 0.10, P < 0.0001, r2 = 0.54; slope = 0.031
± 0.07, P < 0.0001, r2 = 0.94; slope = 0.24 ± 0.06,
P < 0.0001, r2 = 0.94 respectively; Fig 2).
Male and female lionfish differed in their
investment in fat storage (ANCOVA; F1,53 = 3.87,
P = 0.05). Heavier females had lower relative body
fat, while heavier males had higher relative body
fat (females: slope = 0.77 ± 0.44, P = 0.07, r2 =
0.16; males: slope = 1.45 ± 0.26, P < 0.0001, r2 =
0.79; Fig. 3). Gonad mass increased with body
mass in a positive allometric fashion in both sexes
(female: slope = 2.58 ± 0.32, P < 0.0001, r2 =
0.82; male: slope = 1.77 ± 0.17, P < 0.0001, r2 =
0.89; Fig. 4), with females exhibiting a greater rate
of increase than males (F1,53 = 9.85, P = 0.003).
DISCUSSION
Understanding how lionfish invest energy into
growth, reproduction, and defense may be useful
in assessing which factors contribute to their
competitive advantage over native fish species.
Figure 2. Pectoral fin, dorsal spine, and pectoral
spine lengths all had negative allometric
relationships with lionfish weight (slope = 0.20 ±
0.10; slope = 0.031 ± 0.07; slope = 0.24 ± 0.06,
respectively.
133
Little Cayman
Figure 3. Fat mass had a negative allometric
(slope = 0.77 ± 0.44) relationship with weight in
females and a positive allometric (slope = 1.45 ±
0.26) relationship in males.
Figure 4. Gonad mass scaled in a positively
allometric fashion with weight in both females
(slope = 2.58 ± 0.32) and males (slope = 1.77 ±
0.17).
We found that lionfish allocate resources
differently depending on their body size,
ontogenetic stage, and sex. Heavier lionfish appear
to invest relatively less energy into body length
and defensive traits than juveniles (Fig. 2). This
suggests that rapid growth and investment in
predator defense are more important elements to
juvenile survival. Larger lionfish have likely
reached a size threshold at which predation risk is
low.
In both male and female lionfish, gonad
weight and fat storage increase with body mass
(Figs. 3 & 4). However, the allometric scaling of
fat storage and gonad weight with body mass
differed between males and females. As males get
heavier, they exhibit a disproportionate increase in
fat storage (Fig. 3), whereas females invest more
in gonad weight than males at the expense of fat.
The high level of investment in reproduction in
female lionfish has major implications for the
reproductive success of this invasive species in the
Caribbean and central Atlantic. Male and female
lionfish each allocate resources in a way that
confers a competitive advantage against native
fish. Both lionfish exhibit high fecundity and
invest energy disproportionately to reproduction
(Heino & Kaitala 2001). In addition to investment
in greater reproductive capability, males invest in
enhanced fat storage, buffering them against
periods of low food availability and other
environmental stressors (Fishelson 1997). These
life history strategies have the potential to increase
lionfish reproductive success and proliferation,
indicating the management and eradication efforts
should target larger lionfish since they have a
greater reproductive capacity.
LITERATURE CITED
Albins, M.A. and Mark A Hixon. 2013. Worst
case scenario: potential long-term effects of
invasive predatory lionfish (Pterois volitans)
on Atlantic and Caribbean coral-reef
communities. Environmental Biology of
Fishes 96: 115-1157.
Brönmark, C. and J.G. Miner. 1992. Predatorinduced phenotypic change in body
morphology in crucian carp. Science 258:
1348-1350.
Cody, M.L. 1966. A general theory of clutch size.
Evolution 20: 174-184.
Costello, R.A., N.B. Frankel, and M.M. Gamble.
2012. Allometric scaling of morphological
feeding adaptations and extreme sexual
dimorphism in energy allocation in invasive
lionfish (Pterois volitans).
134
ACKNOWLEDGEMENTS
We would like to thank Katie Lohr and Caitlin
Kuempel for outstanding lionfish catching abilities
and unwavering enthusiasm. We would like to
thank Robin Costello, Nina Frankel, and Madeline
Gamble for their project in 2012 and helping make
our project possible. Also, thank you Celia Chen
who sat with us looking at graphs for hours, and
Vivek Venkataraman and Jessica Trout-Haney
who helped us attempt to count gonads.
AUTHOR CONTRIBUTION
All authors contributed equally. Adam Schneider
identified all fish by species.
Dartmouth Studies in Tropical Ecology 2014
Darling, E.S., S.J Green, J.K. O’Leary, and I.M
Cote. 2011. Indo-Pacific lionfish are larger
and more abundant on invaded reefs: a
comparison of Kenyan and Bahamian lionfish
populations. Biological Invasions 13: 20452051.
Fishelson, L. 1997. Experiment and observations
on food consumption, growth, and starvation
in Dendrochirus brachypterus and Pterois
volitans (Pteroinae, Scorpaenidae).
Environmental Biology of Fishes 50: 391-403.
Heino, M. and V. Kaitala. 1999. Evolution of
resource allocation between growth and
reproduction in animals with indeterminate growth.
Journal of Evolutionary Biology 12: 423–429.
Morris, J.A, Jr., C.V. Sullivan, and J.J. Govoni.
2011. Oogenesis and spawn formation in the
invasive lionfish. Scientia Marina 75: 147-154.
135
Little Cayman
IMPACT OF PROTECTED AREAS ON SPINY LOBSTER (PANULIRUS ARGUS)
DISTRIBUTION AND POPULATION STRUCTURE
ALLEGRA M. CONDIOTTE, NATHANAEL J. FRIDAY, AND CLAIRE B. PENDERGRAST
Faculty Editor: Celia Y. Chen
Abstract: Overfishing of economically important tropical reef species has led to widespread declines in fishery yield
and reef ecosystem stability. Populations of Panulirus argus, a highly valued but frequently overexploited top
predator in reef ecosystems, have experienced declines throughout the Caribbean. In response to declining P. argus
abundance in the Cayman Islands, protective regulations such as seasonal closure of fisheries and the creation of
marine protected areas have been implemented. These regulations aim to maintain stable populations of P. argus
inside protected areas and enhance fishery yield by replenishing populations through recruitment from protected to
fished areas. We examined the effect of protection status on P. argus population density, size, and habitat preference
on Little Cayman Island. We found a greater abundance of P. argus in protected versus unprotected areas, but
observed no relationship between body size and protection status. We observed no preference for grass, sand, or
patch reef substrate in lobsters in protected areas, while lobsters in unprotected areas were exclusively located on
patch reefs. Finally, we used our observed population densities in protected and unprotected areas to estimate the
carrying capacity, current population size, and annual harvest of P. argus on Little Cayman Island. Our results
suggest that marine park protection is an effective and essential conservation tactic for maintaining healthy P. argus
populations, but that current populations are unlikely to support increased harvest. Continued management and
enforcement of protective regulations are essential for the health and sustainable harvest of spiny lobster on Little
Cayman Island.
Keywords: marine protected area, fishery management, Panulirus argus, population structure
INTRODUCTION
Management and protection of marine
environments are vital to the stability of ecological
and human systems in coastal areas. Tropical reef
environments are renowned for their biological
diversity and complexity, but they also provide
essential resources and ecosystem services for
coastal communities. As coastal populations have
grown rapidly in recent decades, pressure on
tropical reefs from pollution, development, and
overfishing has increased dramatically. Efforts to
preserve the ecological health and stability of
fishery resources through management of reef
environments have had varied success (Roberts
and Polunin 1993). Economically important
species in tropical reef fisheries are often large
roaming herbivores and carnivores that have a
strong top-down influence on the trophic structure
and community composition of reef ecosystems
(Agardi 2000). Unfortunately, overexploitation
and population decline of highly valued species
occurs in fisheries worldwide, often resulting in
severe losses in both ecological resilience and
economic value of fisheries for coastal
communities (Bellwood et al. 2004).
136
Declining yields from tropical reef fisheries
has resulted in a variety of mitigation efforts,
ranging from total prohibition of fishing activity,
species-specific catch limits and size regulations,
to seasonal closure (Lundquist and Granek 2005).
Marine Protection Areas are of particular
importance in areas where coastal marine tourism
is a central economic driver and where coastal
communities depend on local fisheries for food
and income (Brown et al. 2001). Protected areas
provide a refuge that allows species to rebound
from overexploitation. When they are in close
proximity to unprotected areas, they may also
assist in replenishing unprotected populations by
facilitating movement between these zones (Gell
and Roberts 2003).
Residents of the Cayman Islands have
historically relied on marine systems for food and
income; up until the late 1900s, shipbuilding,
fishing and turtling were primary sources of
income on the island. In recent years a concerted
effort has been underway to promote tourism
development in the Cayman Islands, primarily
through the expansion of ecotourism ventures such
as scuba, snorkel, and fishing activities on the
island’s relatively pristine coral reef ecosystems
Dartmouth Studies in Tropical Ecology 2014
(World Travel Guide, 2014). Spiny lobster is a
popular dish in the Caymans, particularly for
tourists seeking “authentic” Caribbean cuisine
(Marc Pothier, pers. comm). Thus it is likely that
as tourism on the island increases, so will the
demand for this highly valued dish, incentivizing
an increase in lobster harvest beyond current
levels (Conservation). In addition to its economic
value, P. argus is a prominent predator species in
reef ecosystems and plays an important role in the
top-down control of the reef community (Shears
and Babcock 2002). However, the spiny lobster’s
slow growth rate (P. argus estimated age at
maturity is two years and life span is 12 years) and
ease of harvest has resulted in the overexploitation
of populations across the Caribbean (Chavez 2001,
Acosta and Robertson 2003).
Reports of overexploitation and reduced
abundance on Little Cayman Island led to a 1978
Marine Conservation Law imposing P. argus
catch limits, minimum size regulations,
technology restrictions, and closed seasons for
lobster harvest (Cayman Department of
Environment, 2011). In 1985, two replenishment
zones and two marine parks were created,
prohibiting lobster harvest throughout the year in
these areas (Brunt and Davies). As of 2014, 50.7%
of Little Cayman shoreline (19.8 km) is protected
as either Marine Parks, where all marine species
are protected, or as replenishment zones, where
only lobster harvest is prohibited (Bothwell pers.
comm). In the remaining 49.3% of shoreline (19.3
km), lobsters must exceed 6 inches in tail length,
and harvest may not exceed three lobsters per
person or six per boat between 1 December and 28
February (Little Cayman History).
We examined the impact of harvesting activity
on spiny lobster population density, body size, and
habitat preference by surveying P. argus in
protected and unprotected areas on Little Cayman
Island. We hypothesized that P. argus would be
more abundant in protected areas than in
unprotected areas as a result of differences in
fishing pressure. We also predicted a larger mean
body size in the protected area, as larger lobsters
are more susceptible to capture, have greater
economic value, and exhibit increased foraging
activity in exposed environments. Finally, because
they are under greater fishing pressure, we
expected lobsters in unprotected areas to occur
more frequently in the reef where shelter is more
abundant.
METHODS
Data Collection
We surveyed P. argus populations between 2030
and 2330 on 4-6 March 2014 in the Marine Park
unprotected zone of Grape Tree Bay on the north
coast of Little Cayman Island. Sites in the park
and in the unprotected zone were several thousand
meters apart. We surveyed in the marine park on 4
and 5 March, and in the open zone on 5 and 6
March, immediately following the March 1
termination of lobster season.
Figure 1: Representative sampling block, divided into near shore, mid-distance, and back-reef.
137
Little Cayman
We set up twelve sampling blocks: six
continuous blocks in the park and six continuous
blocks in the open zone. Each block extended 50
meters along the coast and 80 meters to the edge
of the back reef (Figure 1). The sampling blocks
crossed three types of substrate: turtle grass, sandy
stretches, and reef. Each sampling block contained
roughly equivalent amounts of turtle grass, sand,
and patch reef substrate. We exhaustively scanned
each block recording abundance of spiny lobsters,
estimated carapace length (Figure 2), and type of
substrate on which the lobster was located. We
searched the perimeter and top of patch reefs but
were unable to exhaustively survey interior
crevices.
Cayman Islands DoE regulations prohibit the
removal of P. argus with tail lengths of less than
Figure 2: Carapace length on P. argus. Tail begins
at the rear end of the carapace.
six inches (152.4 mm) (Cayman Islands
Department of Environment). We used Equation 1
to calculate the minimum carapace length for
harvest (Matthews et al. 2003), roughly
corresponding to the body size at which lobsters
reach maturity (Smith and Herrkind 1992).
Population size, carrying capacity, and annual
harvest estimations
We estimated the overall abundance of P. argus in
the island’s near-shore environments from the
densities observed in our study sites. Protected
areas comprise 19.8 km of the island’s shoreline
and extend roughly 80 meters from shore,
resulting in a protected area of roughly 158.7
hectares. Similarly, 19.3 km of unprotected
shoreline results in an unprotected area of 154.5
hectares. We multiplied each area by the density
of lobsters observed in areas of equivalent
protection status in Grape Tree Bay to provide
rough estimates of population size in Little
Cayman’s protected and unprotected areas
(Equations 2, 3). Adding these two population
estimates provides the total population of P. argus
under current marine protection conditions
(Equation 4). By multiplying the entire island’s
circumference by 80 meters and by the protected
area density, we calculated Little Cayman’s P.
argus carrying capacity (Equation 5). Subtracting
current total population from this carrying capacity
provides an estimate of how many P. argus were
harvested during this past lobster season on Little
Cayman (Equation 6).
(Equation 1)
)(Equation 2)
(Equation 3)
(Equation 4)
(Equation 5)
(Equation 6)
138
Dartmouth Studies in Tropical Ecology 2014
Figure 3. Protection status of Little Cayman Island coastal areas. 50.7% of Little Cayman’s 39 km coastline
is comprised of Marine Park or Replenishment Zone, in which P. argus harvest is prohibited. Protected
survey site was located in the Grape Tree Bay Marine Park, and unprotected site was roughly 1 km east.
Statistical Analyses
We performed an ANOVA comparing density
between sampling blocks within an area to
eliminate the possibility of a block effect in our
results. To examine the difference in population
density between protected and unprotected zones,
we performed a T-test. We performed an ANOVA
and used Tukey’s HSD to compare mean body
size by substrate. Finally, we performed an
ANOVA to compare lobster abundance by
substrate in the protected area.
RESULTS
Lobster density was 3.4 times greater in protected
areas than unprotected areas (T1,30= -2.959, P=
0.006; Figure 4). Lobster size ranged from 4 to 24
cm (carapace length) and did not differ
significantly between protected and unprotected
areas (T1,11= -1.028, P= 0.33). All but five
individuals exceeded the minimum capture size
limit (81.3 mm carapace), indicating that our
Table 1: The mean carapace lengths and
calculated tail lengths for lobsters in the
protected area, in the unprotected area, and
combined.
Figure 4: Density of lobsters per hectare in the
protected area (mean ± SE = 11.25 ± 2.2) and density
in the unprotected area (mean ± SE = 3.33 ± 1.5).
139
Little Cayman
Figure 5: Density of lobsters per hectare by
substrate in the turtle grass (mean ± SE =
13.75 ± 4.1), density in the sand (mean ± SE
= 5.0 ± 2.5), and density in the reef (mean ±
SE = 15.0 ± 3.9).
observations reflected predominantly mature
lobsters (Table 1).
Within the protected area, lobster density did
not differ significantly by substrate (F2,17 = 2.36, P
= 0.13; Figure 5). However, in the unprotected
area, every P. argus we observed was on patch
reef substrate. We found a significant effect of
substrate on carapace length, with smaller lobsters
more frequently on the reef (F2,34= 10.57, P=
0.0003; Figure 6).
We observed a density of 11.25 lobsters per
hectare in the protected area. If Grape Tree Bay is
representative of other protected areas on Little
Cayman, then we estimate that roughly 1750
lobsters are present in protected areas across the
island (Equation 2). The density of lobsters in
unprotected areas (3.33 lobsters per hectare) leads
to an estimated 500 lobsters located in unprotected
areas (Equation 3). The current spiny lobster
population is estimated to be 2300 (Equation 4),
while the carrying capacity for this population is
3524 lobsters (Equation 5). By these estimates,
1223 lobsters, or 35% of carrying capacity, are
removed from unprotected areas during lobster
season each year (Equation 6).
DISCUSSION
140
Figure 6: Carapace length of lobsters by substrate
in the turtle grass (mean ± SE = 166.36 ± 10.8
mm), in the sand (mean ± SE = 182.0 ± 12.4 mm),
and in the reef (mean ± SE = 109.47 ± 10.2 mm).
P. argus is a valuable fishery species in the
Caribbean, where it serves as a major source of
income (Gongora 2010). P. argus were more
abundant in the marine park than in the
unprotected zone (Fig. 4), indicating that the
marine park effectively protects lobster
populations from harvest. P. argus are known to
move up to 200 meters in a night, but rarely leave
their home reefs (Bertelsen and Hornbeck 2009);
considering the distance between the study block
within the park and the unprotected area
(approximately one kilometer), it is unlikely that
migration occurred between the park and the
unprotected zone in the four days between the end
of lobster season and the beginning of our study.
There was no difference in average P. argus
size between the marine park and the unprotected
zone (Table 1). The similarity in size of P. argus
at the two sites suggests that adult lobsters have
not been entirely removed from the unprotected
zone. All but five of the lobsters measured in this
study were above the minimum catch size for
lobster season, with a carapace longer than 813
mm. Spiny lobsters reach maturity between a
carapace length of 76 and 82 mm (Goni et al.
2003), suggesting that the majority of lobsters we
measured were adults. Juvenile P. argus spend
their early lives hiding in macroalgae before
Dartmouth Studies in Tropical Ecology 2014
moving to deep reef dens, making them difficult to
locate. Therefore juveniles may not have been
present at the sites we surveyed, where there was
little macroalgae. Additionally, previous studies
have found no difference in the density of juvenile
spiny lobsters in protected and unprotected areas
(Shears and Babcock 2002). As protective cavities
in patch reefs appeared roughly equivalent in both
sites and minimum size regulations prohibit the
harvest of juveniles, we expect juvenile P. argus
populations to vary little between the two sites.
However, further investigation of distribution and
recruitment of juveniles across protective areas is
essential to determining the stability of Little
Cayman’s P. argus populations.
All of the P. argus in the unprotected area
were found on reef substrate, while lobsters in the
protected area were equally distributed across
grass, sand, and reef substrate (Fig. 5). Lobsters on
the reef are less visible than those in the open, and
may be sheltered from human harvesting during
lobster season. It is possible that the clustering of
unprotected P. argus on the reef is due to pressure
from fishing, as those lobsters who spend more
time hiding in the reef are less likely to be caught
and removed (Wynne and Cote 2007).
While residents of Little Cayman Island no
longer depend on harvest from local waters for
their livelihoods, spiny lobster is a popular dish
among locals and visitors. Sale of P. argus to the
Hungry Iguana Restaurant serves roughly 300
spiny lobster meals annually, generating over
$9,000 CI for local restaurants and over $3,000 CI
for local fishermen (Marc Pothier, pers. comm).
However, should Little Cayman’s population and
tourism industry continue to grow, demand for
spiny lobster may increase, incentivizing increases
in annual harvest that current populations are
likely unable to tolerate.
As a predator species, P. argus is also
fundamental to the shallow reef communities in
which it lives, contributing to top-down regulation
by controlling herbivore populations and
influencing the trophic structure of the reef system
(Shears and Babcock 2002). Overfishing of
lobsters has led to sharp declines in coastal
populations, resulting in the disruption of trophic
cascades and community structure (Acosta and
Robertson 2003). Our results indicate that the
protected status of P. argus in marine parks is an
effective and vital conservation tactic for
maintaining viable lobster populations in the
Caribbean. However, P. argus populations may
not be able to endure increases in harvest beyond
current levels. Continued monitoring and
enforcement of protective regulations is essential
to the ongoing stability of both lobster populations
and their environments on Little Cayman Island.
LITERATURE CITED
Acosta, C. and Robertson, D. 2003. Comparative
spatial ecology of fished spiny lobsters
Panulirus argus and an unfished congener P.
guttatus in an isolated marine reserve at
Glover’s Reef atoll, Belize. Coral Reefs 22:1–
9.
Agardi, T. 2000. Effects of fisheries on marine
ecosystems: a conservationist’s perspective.
ICES Journal of Marine Science 57: 761-765.
Bellwood, D., T. Hughes, C. Folke, and M.
Nyström. 2004. Confronting the coral reef
crisis. Nature. 429: 827-833.
Bertelsen, R.D. and Hornbeck, J. 2009. Using
acoustic tagging to determine adult spiny
lobster (Panulirus argus) movement patterns
in the Western Sambo Ecological Reserve
(Florida, United States). New Zealand Journal
of Marine and Freshwater Research 43: 35-46.
Brown, K. 2001. Trade-off analysis for marine
protected area management. Ecological
Economics 37.3: 417-434.
Brunt, M. and J. Davies, ed. (1994). The Cayman
Islands: Natural History and Biogeography,
Volume 71. Norwell MA: Kluwer Academic
Publishers.
Cayman Islands Department of Environment.
ACKNOWLEDGMENTS
We thank Celia Y. Chen, Jessica Trout-Haney, and
Vivek Venkataraman for their guidance and
assistance in the writing process. We also thank
Amelia Ritger and Aaron Mondshine for their
assistance in data collection.
AUTHOR CONTRIBUTIONS
All authors contributed equally in the data
collection and writing process. Claire B.
Pendergrast developed and designed the project,
and performed all economic modeling.
141
Little Cayman
2000. Investigations into the recruitment,
abundance, and distribution of the Caribbean
spiny lobster, Panulirus argus, into the North
Sound, Grand Cayman, BWI.
Cayman Island Department of the Environment.
2011. Rules for Cayman Islands Marine Parks.
Published online at http://www.doe.ky/wpcontent/uploads/2011/07/Marine-Parks-July2011.pdf, accessed 05 March 2012.
"Conservation." Cayman Islands Legislation.
Web. 08 Mar. 2014. <http://reefresearch.org/
conservation/cayman-islands-conversationlegislation/>.
Gell, F. R., and C. M. Roberts. "Benefits beyond
boundaries: the fishery effects of marine
reserves." Trends in Ecology & Evolution 18.9
(2003): 448-455.
Gongora, M. 2010. Assessment of the spiny
lobster (Panulirus argus) of Belize based on
fishery-dependent data. Marine Fisheries
Institute 4: 5-38.
Lundquist, C. and Granek, E. 2005. Strategies for
Successful Marine Conservation: Integrating
Socioeconomic, Political, and Scientific
Factors. Conservation Biology 19:1771–1778.
"Little Cayman History." LittleCayman.com. Little
Cayman Resort. 8 Mar. 2014. <http://www.
littlecayman.com/our-island/island-history/>.
142
Marx, J.M. and Herrnkind, W.F. 1985.
Macroalgae (Rhodophyta: Laurencia spp.) as
habitat for young juvenile spiny lobsters,
Panulirus argus. Bulletin of Marine Science
36: 423-431.
Matthews, T.R., Hunt, J.H., and Heatwole, D.W.
2003. Morphometrics and management of the
Caribbean spiny lobster, Panulirus argus.
Proceedings of the Gulf and Caribbean
Fisheries Institute 54:156-174.
Shears, N.T. and Babcock, R.C. 2002. Marine
reserves demonstrate top-down control of
community structure on temperate reefs.
Oecologia 132:131-142.
Roberts, C. M., and N. V. C. Polunin. 1993.
Marine reserves: Simple solutions to
managing complex fisheries? Ambio 22: 363368.
Smith, K.N., and Herrkind, W.F. 1992. Predation
on early juvenile spiny lobsters Panulirus
argus (Latreille): influence of size and shelter.
Journal of Experimental Marine Biology and
Ecology 157: 3-18.
Wynne, S. and Cote, L. 2007. Effects of habitat
quality and fishing on Caribbean spotted spiny
lobster populations. Journal of Applied
Ecology. 44: 488-494.
Dartmouth Studies in Tropical Ecology 2014
MY ANEMONE, MY RULES: NICHE PARTITIONING IN ANEMONE COMMUNITIES
LARS-OLAF HÖGER, ELLEN R. PLANE, AND ADAM N. SCHNEIDER
Faculty Editor: Celia Y. Chen
Abstract: Coral reefs are known for their immense species diversity. However, many coral reef species are
functionally redundant, suggesting that high diversity is maintained through mechanisms that reduce competition.
We studied functionally similar shrimp and crabs inhabiting Bartholomea annulata and Stichodactyla helianthus
anemones at Little Cayman Island to determine how these crustaceans are able to coexist. We found that anemone
size does not have an effect on the abundance of inhabitants, suggesting that space is not a limiting factor for
anemone inhabitants. We also found that species in the same functional feeding group never inhabit the same
anemone, indicating that competition between inhabitants of the same functional feeding group limits the
distributions of species. We conclude that spatial niche partitioning plays a key role in maintaining the diversity of
coral reefs.
Key words: anemone, functional feeding groups, niche partitioning, shrimp
INTRODUCTION
Coral reefs are widely recognized for their
immense species diversity. Within a reef
ecosystem, many animal species fill the same or
similar functional niches (Bellwood et al. 2003),
suggesting that mechanisms are operating to
reduce competition. Different species of anemone
shrimp, which inhabit a variety of anemone
species and live amongst the tentacles for
protection, often coexist and appear to fill similar
feeding and protective niches (Smith 1977, Wirtz
1997). Anemones host a number of cleaner shrimp
species, as well as scavengers and predators. For
example, both Periclimenes yucatanicus and
Ancylomenes pedersoni are cleaner shrimp that
commonly inhabit Bartholomea annulata
anemones (Silbiger and Childress 2008).
Competition between inhabitant species for
protective space, food scraps, or cleaner clients
may limit the number of inhabitant species that
can successfully reside on one anemone.
We conducted an observational study of
inhabitant species of the anemones Bartholomea
annulata (corkscrew anemone) and Stichodactyla
helianthus (sun anemone) to examine the factors
affecting species distribution and coexistence
within a reef ecosystem. We expected that larger
anemones would host more shrimp because they
have more usable surface area and thus a higher
carrying capacity. These anemones should host
more cleaner shrimp in particular, as larger
anemones are more visible signals to fish that
require cleaning services (Chadwick and Huebner,
2012). We also hypothesized that anemone
inhabitants that fill the same functional feeding
niche would exhibit spatial niche partitioning to
avoid direct competition. We predicted that shrimp
of different functional feeding groups (i.e.
scavenger and cleaner) would be found coexisting
on the same anemone, but multiple shrimp species
of the same functional group (i.e. cleaner and
cleaner) would not. It is also possible that
inhabitant species may specialize to specific
species of anemone, which would fulfill the same
niche partitioning function on a species-wide
rather than individual anemone level. Both of
these methods of niche partitioning should reduce
interspecific competition for shelter and food.
METHODS
We haphazardly sampled 15 corkscrew anemones
and 28 sun anemones in Grape Tree Bay and
South Hole Sound, Little Cayman, on 5 and 6
March 2014. We counted anemones that were in
the process of splitting, or had joined discs, as a
single anemone. We measured the length across
the longest axis for each anemone disc and noted
the substrate that it was growing on (rock, sand,
seagrass, coral, or conch). We then counted the
number of shrimp of each species on the anemone,
and recorded the number of crabs (unidentified
species) present. Since many anemones were
surrounded by hermit crabs, we also recorded
whether or not hermit crabs were present.
143
Little Cayman
RESULTS
We observed four shrimp species (Alpheus sp.,
Periclimenes rathbunae, Thor amboinensis, and
Palaemonetes sp.), in the two species of
anemone we studied, as well as several
unidentified crab species. Alpheus sp. is a
predatory species, P. rathbunae a cleaner
species, and T. amboinensis and Palaemonetes
sp. are scavenger species (Humann and
DeLoach 1992).
Anemone size did not have an effect on
number of inhabitants (Fig. 1), or number of P.
rathbunae, the only cleaner shrimp species we
observed. We tested for a size effect only in sun
anemones because corkscrew anemones hosted
a maximum of two inhabitants. Substrate type
did not have an effect on the abundance of any
inhabitant species.
We observed that 73% of corkscrew anemones
and 96% of sun anemones had at least one shrimp
or crab inhabitant. Alpheus sp. was only present
in corkscrew anemones, and P. rathbunae and T.
amboinensis were only present in sun anemones.
Palaemonetes sp. and crabs were most
frequently in sun anemones, but also
occasionally in corkscrew anemones.
We found each species alone on at least one
anemone. Alpheus sp. was the sole inhabitant of
an anemone more frequently than any other
species (Fig. 2). Crabs and P. rathbunae
coexisted on the same anemone more frequently
than any other two species (Fig. 2) and we never
saw T. amboinensis and Palaemonetes sp. on the
same anemone. The presence of hermit crabs did
not have an effect on abundance of any other
inhabitant species.
The first two principal components (PC1
and PC2) accounted for 36% and 22 % of the
144
variance, respectively. According to the first
component of a PCA on community
composition, Alpheus sp. was unlikely to be
associated with any other inhabitant species.
The second component showed that P.
rathbunae, Palaemonetes sp., and crabs were
likely to be associated with one another, but not
with T. amboinensis (Table 1, Fig. 3).
Table 1. Loading matrix for first two principal
components of community structure.
PC1
PC2
P. rathbunae
0.75
-0.07
Alpheus sp.
-0.73
-0.21
T. amboinensis
0.16
0.89
Palaemonetes sp.
0.33
-0.47
Crab
0.77
-0.12
12
Number of inhabitants
Data analyses
We performed linear regressions between
anemone disc length and number of anemone
inhabitants and number of cleaner shrimp for each
species of anemone. We also conducted a
Principal Components Analysis (PCA) on the
abundances of each species to obtain a measure of
community composition. Finally, we conducted
analysis of variance (ANOVA) to examine the
effect of substrate on the abundance of inhabitants.
10
8
6
4
2
0
0
10
20
30
Anemone length (cm)
Figure 1. Anemone length did not have an effect on the
number of inhabitants.
DISCUSSION
Our results indicate that spatial niche partitioning
between functionally similar species is playing a
role in maintaining diversity on coral reefs. The
two species of scavenger shrimp were never found
on the same individual anemone, suggesting that
they may be actively separating based on
competition within their functional feeding group.
This example of spatial niche partitioning may
help to explain how direct competition can be
avoided within the reef by choosing different
residence locations to avoid overlap. The frequent
Dartmouth Studies in Tropical Ecology 2014
association of crabs with cleaner shrimp supports
this theory as well; many crabs are scavengers and
opportunistic hunters (Johnston and Freeman
2005), so their coexistence with non-scavenging
shrimp suggests further niche partitioning. Spatial
niche partitioning may also occur within the
anemone as all crabs were found sheltering
underneath the disc rather than in the tentacles
with the shrimp. Although niche partitioning may
be occurring in cleaner and predatory functional
groups, we were unable to observe it because our
study only included one species in each of these
groups.
10
9
Number of Anemones
8
7
6
5
4
3
2
1
0
Figure 2. Alpheus sp. was frequently found alone, while other species were often found coexisting with other species
on the same anemone. Black bars are single-species assemblages, white bars two-species assemblages, and
hatched bars three-species assemblages. Only anemones with inhabitants were included.
Our results demonstrate that species in the
same functional feeding group are generally found
on separate individual anemones, but not on
different species of anemones. The two scavenger
species both showed a preference for sun
anemones, athough one Palaemonetes sp. shrimp
was found on a corkscrew anemone, indicating
that it is also a viable habitat for these shrimp.
Alpheus sp. showed a strong preference for
corkscrew anemones and no Alpheus sp. shrimp
were found on sun anemones. Corkscrew
anemones have longer tentacles (Humann and
DeLoach 1992), which may be more adequate for
sheltering the large Alpheus sp. than the smaller
sun anemone tentacles. We never saw Alpheus sp.
coexisting with other shrimp species. This is most
likely due to their predatory nature; Alpheus sp.
hunt smaller shrimp and crabs and would not be
145
Little Cayman
expected to share anemone disc space with
potential prey.
Since almost all of the anemones sampled
contained at least one shrimp or crab, it is likely
that these anemones are a limiting resource.
However, space within the anemone is not limited;
in sun anemones, inhabitant abundance was not
affected by anemone size, which suggests that
competition for anemone protection among
inhabitants is not density-dependent. Anemone
size also had no effect on cleaner shrimp
abundance, indicating that larger anemones are not
a more desirable resource for cleaner shrimp. It is
possible that signaling client fish from a distance
is irrelevant to the shrimp on these particular
anemones because they were found at shallow
depths, or that the shrimp are relying on other
features of the anemone, such as color or shape, to
signal client fish. Substrate did not have an effect
on anemone community composition, suggesting
that the anemone itself, rather than the
environment it lives in, has the greatest effect on
the species that inhabit it.
Niche partitioning in the scavenger functional
group could be an indication of what is occurring
in the rest of the reef ecosystem, as other animals
may also be separating spatially based on feeding
strategy. This partitioning does not seem to be
affected by anemone size or abundance of
anemone inhabitants, but rather by an avoidance of
overlap between identical functional groups. This
mechanism of species coexistence, employed as a
strategy to avoid competition, could help explain
how such a high level of diversity is maintained
within coral reefs.
LITERATURE CITED
Huebner, L.K., and N.E. Chadwick. 2012. Reef
fishes use sea anemones as visual cues for
cleaning interactions with shrimp. Journal
Experimental Marine Biology and Ecology
416−417: 237−242.
Humann, P., and N. DeLoach. 1992. Reef
Creature Identification: Florida, Caribbean,
Bahamas. New World Publications, Inc.,
Jacksonville, Florida. Page 93.
Johnson, D., Freeman, J. 2005. Dietary
preference and digestive enzyme activities as
indicators of trophic resource utilization by six
species of crab. The Biological Bulletin.
208:36-46
Silbiger, N.J., and M.J. Childress.
2008. Interspecific Variation in Anemone
Shrimp Distribution and Host Selection in the
Florida Keys (USA): Implications for Marine
Conservation. Bulletin of Marine Science
83: 329-345. Rican landscapes.
Smith, W. L. 1977. Beneficial behavior of a
symbiotic shrimp to its host anemone. Bulletin
of Marine Science 27: 343–346.
Dartmouth Studies in Tropical Ecology 2012, in
press.
Wirtz, P. 1997. Crustacean symbionts of the sea
anemone Telmatactis cricoides at Madeira and
the Canary Islands. Journal of Zoology 242:
799–811.
146
Figure 3. The first principal component of a PCA
showed that Alpheus sp. was unlikely to cohabitate with
any other anemone inhabitant species, and the second
principal component showed that T. amboinensis was
unlikely to cohabitate with P. rathbunae, crabs, or
Palaemonetes sp.
ACKNOWLEDGEMENTS
We would like to thank Celia Chen, Jessica TroutHaney, and Vivek Ventakaraman for their
guidance in assembling this paper.
AUTHOR CONTRIBUTIONS
All authors contributed equally.