0% found this document useful (0 votes)
81 views496 pages

EB 2742 kNJ2kaf1

Uploaded by

adritaabbas
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
81 views496 pages

EB 2742 kNJ2kaf1

Uploaded by

adritaabbas
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 496

An Introduction to

Conservation Biology
An Introduction to

Conservation
Biology

Richard B. Primack
Boston University

Anna A. Sher
University of Denver

Sinauer Associates, Inc. Publishers


Sunderland, MA U.S.A.
The Cover
The golden lion tamarin (Leontopithecus rosalia) is an endangered
species of Brazil’s Atlantic Forest that had declined dramatically in
the wild. Strong conservation action involving protecting the re-
maining populations and their habitat, developing captive breed-
ing colonies, and establishing new populations in the wild have
allowed the species to recover from the brink of extinction.
(© Juan Carlos Muñoz/AGE Fotostock.)

An Introduction to Conservation Biology


Copyright © 2016 by Sinauer Associates, Inc.
All rights reserved. This book may not be reproduced in whole or in part
without permission from the publisher. For information or to order, address:
Sinauer Associates
P.O. Box 407
Sunderland, MA 01375 U.S.A.
E-mail: orders.sinauer.com; [email protected]
Internet: www.sinauer.com

Library of Congress Cataloging-in-Publication Data

Names: Primack, Richard B., 1950- author. | Sher, Anna, author.


Title: Introduction to conservation biology / Richard B. Primack, Anna A.
Sher.
Description: Massachusetts, U.S.A. : Sinauer Associates, Inc. Publishers,
2016. | Includes bibliographical references and index.
Identifiers: LCCN 2016005729 | ISBN 9781605354736 (pbk. : alk. paper)
Subjects: LCSH: Conservation biology.
Classification: LCC QH75 .P7523 2016 | DDC 333.95/16--dc23
LC record available at http://lccn.loc.gov/2016005729

Printed in U.S.A.
5 4 3 2 1
We dedicate this book to our students, and to all those
who expand our knowledge of biodiversity and its
protection by engaging in conservation biology.
Contents
Chapter 1
Defining Conservation Biology 2
The New Science of Conservation Biology 5
The roots of conservation biology 6
A new science is born 12
The interdisciplinary approach: A case study
with sea turtles 13
The Ethical Principles of Conservation
Biology 15
Looking to the Future 17
Summary 18

Chapter 2
What Is Biodiversity? 22
Species Diversity 25
What is a species? 25
Measuring species diversity 28
Genetic Diversity 31
Ecosystem Diversity 33
What are communities and ecosystems? 33
Species interactions within ecosystems 35
Food chains and food webs 37
Keystone species and resources 38
Ecosystem dynamics 41
viii Contents

Biodiversity Worldwide 41
How many species exist worldwide? 41
Where is the world’s biodiversity found? 45
The distribution of species 47
Summary 48

Chapter 3
The Value of Biodiversity 52
Ecological and Environmental Economics 55
Cost–benefit analysis 57
Financing conservation 59
What are species worth? 60
Ecosystem services 60
Economic Use Values 61
Direct use values 61
Consumptive use value 61
Productive use value 63
Indirect use values 68
Ecosystem productivity 68
Water and soil protection 70
Climate regulation 72
Species relationships and environmental
monitors 73
Amenity value 74
Educational and scientific value 77
Multiple uses of a single resource:
A case study 77
The Long-Term View: Option Value 78
Existence Value 80
Environmental Ethics 83
Ethical values of biodiversity 83
Deep ecology 86
Summary 87
Contents ix

Chapter 4
Threats to Biodiversity 90
Human Population Growth and Its Impact 92
Habitat Destruction 96
Tropical rain forests 99
Other threatened habitats 102
Desertification 105
Habitat Fragmentation 106
Threats posed by habitat fragmentation 107
Edge effects 110
Environmental Degradation and Pollution 112
Pesticide pollution 113
Water pollution 114
Air pollution 117
Global Climate Change 118
Ocean acidification, warming, and rising sea
level 122
The overall effect of global warming 124
Overexploitation 126
International wildlife trade 128
Commercial harvesting 130
Invasive Species 132
Threats posed by invasive species 134
Invasive species on oceanic islands 135
Invasive species in aquatic habitats 136
The ability of species to become invasive 137
Control of invasive species 140
GMOs and conservation biology 141
Disease 142
A Concluding Remark 146
Summary 146
x Contents

Chapter 5
Extinction Is Forever 150
The Meaning of “Extinct” 154
The current, human-caused mass
extinction 155
Local extinctions 158
Extinction rates in aquatic environments 159
Measuring Extinction 160
Background extinction rates 161
Extinction rate predictions and the island
biogeography model 161
Extinction rates and habitat loss 164
Vulnerability to Extinction 166
Problems of Small Populations 172
Loss of genetic diversity 172
Consequences of reduced genetic diversity 176
Factors that determine effective population
size 179
Other factors that affect the persistence of small
populations 184
Demographic stochasticity 184
Environmental stochasticity and
catastrophes 187
The extinction vortex 188
Summary 189

Chapter 6
Conserving Populations and Species 192
Applied Population Biology 194
Methods for studying populations 195
Population viability analysis (PVA) 203
Metapopulations 208
Long-term monitoring 210
Conservation Categories 212
Prioritization: What Should Be Protected? 216
Contents xi

Legal Protection of Species 222


National laws 222
International agreements to protect species 228
Summary 232

Chapter 7
Bringing Species Back from the Brink 234
Establishing and Reinforcing Populations 236
Considerations for animal programs 238
Behavioral ecology of released animals 242
Establishing plant populations 243
The status of new populations 244
Ex Situ Conservation Strategies 246
Zoos 249
Aquariums 255
Botanical gardens 256
Seed banks 258
Can Technology Bring Back Extinct
Species? 261
Summary 262

Chapter 8
Protected Areas 264
Establishment and Classification of Protected
Areas 266
Marine protected areas (MPAs) 271
The effectiveness of protected areas 272
Measuring effectiveness: Gap analysis 274
Designing Protected Areas 277
Protected area size and characteristics 280
Networks of Protected Areas 283
Habitat corridors 284
xii Contents

Landscape Ecology and Park Design 286


Managing Protected Areas 288
Managing sites 290
Monitoring sites 292
Management and people 294
Zoning as a solution to conflicting
demands 294
Biosphere reserves 296
Challenges to Park Management 298
Poaching 298
Trophy hunting 298
Human–animal conflict 299
Degradation 299
Climate change 299
Funding and personnel 300
Summary 301

Chapter 9
Conservation Outside Protected Areas 304
The Value of Unprotected Habitat 307
Military land 308
Unprotected forests 310
Unprotected grasslands 310
Unprotected waters 311
Land that is undesirable to humans 312
Private land 312
Conservation in Urban and Other
Human-Dominated Areas 313
Other human-dominated landscapes 315
Ecosystem Management 318
Working with Local People 321
Biosphere reserves 324
In situ agricultural conservation 325
Extractive reserves 325
Community-based initiatives 327
Contents xiii

Payments for ecosystem services 328


Evaluating conservation initiatives that involve
traditional societies 330
Case Studies: Namibia and Kenya 330
Summary 334

Chapter 10
Restoration Ecology 336
Where to Start? 339
Restoration in Urban Areas 344
Restoration Using Organisms 346
Moving Targets of Restoration 350
Restoration of Some Major Communities 351
Wetlands 351
Aquatic systems 353
Prairies and farmlands 355
Tropical dry forest in Costa Rica 357
The Future of Restoration Ecology 359
Summary 360

Chapter 11
The Challenges of Sustainable
Development 362
Sustainable Development at the Local
Level 365
Local and regional conservation
regulations 365
Land trusts and related strategies 367
Enforcement and public benefits 371
Conservation at the National Level 372
International Approaches to Sustainable
Development 374
International Earth summits 374
International agreements that protect
habitat 379
xiv Contents

Funding for Conservation 383


The World Bank and international
nongovernmental organizations
(NGOs) 383
Environmental trust funds 386
Debt-for-nature swaps 387
How effective is conservation funding? 387
Summary 389

Chapter 12
An Agenda for the Future 392
Ongoing Problems and Possible Solutions 394
The Role of Conservation Biologists 402
Challenges for conservation biologists 402
Achieving the agenda 403
Summary 408

Appendix: Selected Environmental


Organizations and Sources of
Information 411
Glossary 415
Chapter Opening Photograph
Credits 424
Bibliography 425
Index 463
About the Author 477
Preface
The field of Conservation Biology is in a period of rapid growth, collabo-
ration, increased clarity, and optimism. In the last two years, the success
of intensive conservation efforts have been seen in the growing popula-
tions of wolves and bears in Yellowstone National Park, the establishment
of the largest contiguous ocean reserve to date (830,000 km2) by the U.K.,
and the discovery of hundreds of new plants and animals across the globe,
to only name a few. The public is increasingly engaged with the news of
conservation biology, as reflected in the outrage over the 2015 poaching
death of Cecil the Lion in Zimbabwe. It can also be seen in the arts, such as
the excitement surrounding the 2016 release of Finding Dory, an animated
movie with strong conservation themes, and by the powerful documentary
about orcas, Blackfish, which spurred national legislation in 2014 and major
changes by SeaWorld in 2016 that support the species. However, the task
before us is still vast; as the recent update to the IUCN’s list of endangered
species has shown, most groups of organisms face greater extinction risk
today than ever before as the global human population reaches above 7.25
billion. But if the 2015 Paris Accord on climate change is any indication,
the human race is also more united and serious than ever before about
confronting these challenges. It is our hope that this book can be an im-
portant tool for the development of the next generation.
An Introduction to Conservation Biology is the heir of our two previous,
successful volumes: A Primer of Conservation Biology, Fifth Edition and Es-
sentials of Conservation Biology, Sixth Edition. Since having dedicated myself
to this project, I now respect—at an entirely new level—what Richard
Primack has done, understanding not only the hours involved but the
care it takes to keep content current, coverage across ecosystems and taxa
balanced, and to support the text with images and tables that are beauti-
ful, useful, and themselves representative of the diversity of the field at
all levels.
My goals have to been to retain the readability of A Primer of Conserva-
tion Biology with the depth and coverage of Essentials of Conservation Biology
while upholding the standards of each book to the highest degree. For those
who have used either of these texts, Introduction should feel both familiar
and fresh: three chapters have been added to those offered in the Primer,
with population biology, conservation tools, restoration ecology, sustain-
able development, ex situ conservation, and other key elements expanded
xvi Preface

and updated. There are literally hundreds of new examples, explanations,


citations, and figures to enhance learning and excitement for the subject. It
is designed to serve a wide variety of courses and students, as a primary
textbook or as a supplement.
As Richard has said,
It is our goal that readers of this book will be inspired to find out more
about the extinction crisis facing species and ecosystems and how they
can take action to halt it. I encourage readers to take the field’s activist
spirit to heart—use the Appendix to find organizations and sources of
information on how to help. If readers gain a greater appreciation for
the goals, methods, and importance of conservation biology, and if they
are moved to make a difference in their everyday lives, this textbook will
have served its purpose.

Acknowledgments
I sincerely appreciate the contributions of everyone who has made this
textbook what it is today. Individual chapters were reviewed by J. Michael
Reed, Andrew R. Blaustein, Dana Bauer, John J. Cox, Scott Connelly, Mi-
chael Reed, Peter Houlihan, Tim Caro, Paige Warren, Janette Wallis, Mo-
nique Poulin, Eric Higgs, Federico Cheever, Meg Lowman, and Richard
Reading (who also provided several images).
I am grateful for the work of Annie Henry, Eduardo González, and
Robert Robinson for their background research for the book and helping
to edit material from Essentials into the leaner format with new citations
and examples. Additional help was provided by Allison Brunner, Brandon
Krentz, Jaime Pena, and Matt Herbert, and I appreciate the donation of
photos from many fine photographers, including Scott Dressel-Martin of
Denver Botanic Gardens, Wright Robinson, and Hector R. Chenge. Andy
Sinauer, Rachel Meyers, Martha Lorantos, David McIntyre, Ann Chiara
and the rest of the Sinauer staff helped to transform the manuscript into
a finished book; I greatly appreciated their encouragement and patience
over the past several months.
I would also like to give special thanks to my wife, Fran, and our son
Jeremy as well as all the women in my reading and writing groups who
have offered support and encouragement along the way. Thanks to my
chair, Joseph Angelson, and the Organismal Biologists group at the Univer-
sity of Denver, for their shared enthusiasm for this project. I am indebted
to my hundreds of students who have taught me so much about the art
and science of teaching.
Finally, I must thank Richard Primack for the honor of being his suc-
cessor, for his mentorship, and for all of his input at every stage. It has been
a deeply rewarding experience.
Anna A. Sher
Denver, Colorado
April 6, 2016
Preface xvii

An International Approach
In keeping with the global nature of conservation biology, I feel it is impor-
tant to make the field accessible to as wide an audience as possible. With
the assistance of Marie Scavotto and the staff of Sinauer Associates, I have
arranged an active translation program of the Essentials of Conservation
Biology and the shorter Primer of Conservation Biology. The goal has been to
create regional or country-specific translations, identifying local scientists
to become coauthors and to add case studies, examples, and illustrations
from their own countries and regions that would be more relevant to the
intended audience. Already, editions of have appeared in Arabic, Brazilian
Portuguese, Chinese (four editions), Czech (two editions), Estonian, French
(two editions, one with a Madagascar focus), German, Greek, Hungarian,
Indonesian (two editions), Italian (two editions), Japanese (two editions),
Korean (three editions), Mongolian, Nepal (in English), Romanian (two
editions), Russian, Serbia, South Asia (in English), Spanish (two editions;
one with a Latin American focus), Turkish, and Vietnamese. New editions
for Africa, Bangladesh, Brazil, Germany, Greece, Iran, Laos, Latin America,
Madagascar, Pakistan, and Thailand are currently in production. These
translations will help conservation biology develop as a discipline with a
global scope. At the same time, examples from these translations find their
way back into the English language editions, thereby enriching the presen-
tation. I hope that my enthusiastic new coauthor, Anna Sher, will continue
this project with our new book, An Introduction to Conservation Biology.

Richard Primack
Boston University
April 6, 2016
Media and Supplements
to accompany

An Introduction to
Conservation Biology
eBook
An Introduction to Conservation Biology is available for purchase as an eBook,
in several different formats, including VitalSource, Yuzu, BryteWave, and
RedShelf. The eBook can be purchased as either a 180-day rental or a per-
manent (non-expiring) subscription. All major mobile devices are support-
ed. For details on the eBook platforms offered, please visit www.sinauer.
com/ebooks.

Instructor’s Resource Library


(Available to qualified adopters)
The Instructor’s Resource Library for An Introduction to Conservation Biology
includes all of the textbook’s figures (line-art illustrations and photographs)
and tables, provided as both high- and low-resolution JPEGs. All have been
formatted and optimized for excellent projection quality. Also included
are ready-to-use PowerPoint slides of all figures and tables.
An Introduction to
Conservation Biology
Defining
1 Conservation Biology
The New Science of
Conservation Biology 5
The Ethical Principles of
Conservation Biology 15
Looking to the Future 17

A conservation biologist from the binational


Gladys Porter Zoo crew releases Kemp’s ridley
sea turtles off the coast of Tamaulipas, Mexico.
P
opular interest in protecting the world’s bio-
logical diversity—including its amazing range
of species, its complex ecosystems, and the
genetic variation within species—has intensified dur-
ing the last few decades. It has become increasingly
evident to both scientists and the general public that
we are living in a period of unprecedented losses of
biodiversity* (Pimm et al. 2014). Around the globe,
biological ecosystems—the interacting assemblages
of living organisms and their environment that took
millions of years to develop—including tropical rain
forests, coral reefs, temperate old-growth forests,
and prairies, are being devastated. Thousands, if not
tens of thousands, of species and millions of unique
populations are predicted to go extinct in the coming
decades (Barnosky et al. 2011). Unlike the mass ex-
tinctions in the geologic past, which followed catas-
trophes such as asteroid collisions with Earth, today’s
extinctions have a human face.
During the last 200 years, the human population
has exploded. It took more than 160,000 years for the
number of Homo sapiens to reach 1 billion, an event

*Biological diversity is often shortened to biodiversity (a term


credited to biologist E. O. Wilson in 1992); it includes all species,
genetic variation, and biological communities and their ecosys-
tem-level interactions.
4 Chapter 1

FIGURE 1.1 The human population in 10


2016 stood at about 7.25 billion. The World 9.4
Resources Institute estimates current an- 9
nual population growth at 1.1%, but even
this modest growth rate will add more 8
than 70 million people to the planet in the
next year. This number will escalate each
7

Human population (billions)


year as the increase is compounded. (Data
from US Census Bureau, www.census.gov) 6.1
6

3
2.5

2
1.7
1.2
1

0
1850 1900 1950 2000 2050
(estimated)
Year

that occurred sometime around the year 1805. Estimates for 2016 put the
number of humans at 7.25 billion, with a projected 9.4 billion by 2050 (US
Census Bureau); at this size, even a modest rate of population increase
adds tens of millions of individuals each year (Figure 1.1). The threats
to biodiversity are accelerating because of the demands of the rapidly in-
creasing human population and its rising material consumption. People
deplete natural resources such as firewood, coal, oil, timber, fish, and game,
and they convert natural habitats to land dominated by agriculture, cit-
ies, housing developments, logging, mining, industrial plants, and other
human activities. These changes are not easily reversible, and even ag-
gressive programs to slow population growth do not adequately address
the environmental problems
Introduction towe have caused
Conservation Biology(Bradshaw and Brook 2014).
1E Primack/Sher
Sinauer Associates
Worsening the situation is the fact that as countries develop and indus-
Morales Studio/SA
trialize, the consumption of resources increases.
Primack_Sher1E_01.01 Date For example,
10-29-15 the average
3-3-16
citizen of the United States uses four times more energy than the average
global citizen, seven times more than the average Costa Rican citizen, and
fifteen times more than the average Indian citizen (US Energy Informa-
tion Administration 2015). The ever-increasing number of human beings
and their intensifying use of natural resources have direct and harmful
consequences for the diversity of the living world (Brown et al. 2014).
Defining Conservation Biology 5

The threats to biodiversity directly threaten human populations as well


because people are dependent on the natural environment for raw materi-
als, food, medicines, and the water they drink. The poorest people are,
of course, the ones who experience the greatest hardship from damaged
environments because they have fewer reserves of food and less access to
medical supplies, transportation, and construction materials.

The New Science of Conservation Biology


The avalanche of species extinctions and the wholesale habitat destruction
occurring in the world today is devastating, but there is reason for hope.
The last several decades have included many success stories, such as that
of the American bald eagle (Haliaeetus leucocephalus), which was rescued
from near extinction due to a combination of scientific inquiry, public
awareness, and political intervention (see Chapter 11). Actions taken—or
bypassed—during the next few decades will determine how many of the
world’s species and natural areas will survive. It is quite likely that people
will someday look back on the first decades of the twenty-first century as
an extraordinarily exciting time, when collaborations of determined peo-
ple acting locally and internationally saved many species and even entire
ecosystems (Sodhi et al. 2011). Examples of such conservation efforts and
positive outcomes are described throughout this book.
Conservation biology is an integrated, multidisciplinary scientific field
that has developed in response to the challenge of preserving species and
ecosystems. It has three goals:
• To document the full range of biological diversity on Earth
• To investigate human impact on species, genetic variation, and
ecosystems
• To develop practical approaches to prevent the extinction of species,
maintain genetic diversity within species, and protect and restore bio-
logical communities and their associated ecosystem functions

The first two of these goals involve the dispassionate search for factual
knowledge that is typical of scientific research. The third goal, however,
defines conservation biology as a normative discipline —that is, a field
that embraces certain values and attempts to apply scientific methods to
achieving those values (Lindenmayer and Hunter 2010). Just as medical
science values the preservation of life and health, conservation biology
values the preservation of species and ecosystems as an ultimate good, and
its practitioners intervene to prevent human-caused losses of biodiversity.
Conservation biology arose in the 1980s, when it became clear that the
traditional applied disciplines of resource management alone were not
comprehensive enough to address the critical threats to biological diversity.
The applied disciplines of agriculture, forestry, wildlife management, and
6 Chapter 1

fisheries biology have gradually expanded to include a broader


Conservation biology range of species and ecosystem processes. Conservation biology
merges applied and complements those applied disciplines and provides a more gen-
theoretical biology by eral theoretical approach to the protection of biological diversity.
incorporating ideas It differs from these disciplines in its primary goal of long-term
and expertise from a preservation of biodiversity.
broad range of scientific Like medicine, which applies knowledge gleaned from
fields toward the goal of physiology, anatomy, biochemistry, and genetics to the goal of
achieving human health and eliminating illness, conservation
preserving biodiversity.
biology draws on other academic disciplines, including popula-
tion biology, taxonomy, ecology, and genetics. Many conserva-
tion biologists have come from these ranks. Others come from
backgrounds in the applied disciplines, such as forestry and wildlife man-
agement. In addition, many leaders in conservation biology have come from
zoos and botanical gardens, bringing with them experience in locating rare
and endangered species in the wild and then maintaining and propagating
them in captivity.
Conservation biology is also closely associated with, but distinct from,
environmentalism, a widespread movement characterized by political and
educational activism with the goal of protecting the natural environment.
Conservation biology is a scientific discipline based on biological research
whose findings often contribute to the environmental movement (Hall
and Fleishman 2010).
Because much of the biodiversity crisis arises from human pressures,
conservation biology also incorporates ideas and expertise from a broad
range of fields outside of biology (Figure 1.2) (Reyers et al. 2010). For ex-
ample, environmental law and policy provide the basis for government pro-
tection of rare and endangered species and critical habitats. Environmental
ethics provides a rationale for preserving species. Ecological economists
provide analyses of the economic value of biological diversity to support
arguments for preservation. Climatologists monitor the physical charac-
teristics of the environment and develop models to predict environmental
responses to disturbance. Both physical and cultural geography provide
information about the relationships among elements of the environment,
helping us understand causes and distributions of biodiversity and how
humans interact with it. Social sciences, such as anthropology and sociol-
ogy, provide methods to involve local people in actions to protect their
immediate environment. Conservation education links academic study
and fieldwork to solve environmental problems, teaching people about
science and helping them realize the value of the natural environment.
Because conservation biology draws on the ideas and skills of so many
separate fields, it can be considered a truly multidisciplinary discipline.

The roots of conservation biology


Religious and philosophical beliefs about the relationship between humans
and the natural world are seen by many as the foundation of conserva-
Defining Conservation Biology 7

FIGURE 1.2 Conservation


rience and research n
ld expe eed biology represents a synthe-
Fie s
sis of many basic sciences
(left) that provide principles
and new approaches for the
Basic Sciences Resource Management applied fields of resource
Anthropology Agriculture management (right). The ex-
Biogeography Community education periences gained in the field,
Climatology and development
Ecology: Fisheries management in turn, influence the direction
Community ecology Forestry of the basic sciences. (After
Ecosystem ecology Land-use planning and Temple 1991.)
Landscape ecology regulation
Environmental studies: Management of captive
Ecological economics populations:
Environmental ethics Zoos
Environmental law Aquariums
Ethnobotany Botanical gardens
Evolutionary biology Seed banks
Genetics Management of protected
Population biology areas
Sociology Sustainable development
Taxonomy Wildlife management
Other biological, physical, Other resource conservation
and social sciences and management activities

New
id eas and approaches

tion biology (Dudley et al. 2009). Eastern philosophies such as Taoism,


Hinduism, and Buddhism revere wilderness for its capacity to provide
intense spiritual experiences. These traditions see a direct connection be-
tween the natural world and the spiritual world, a connection that breaks
down when the natural world is altered or destroyed. Strict adherents to
the Jain and Hindu religions in India believe that all killing of animal life
is wrong. Islamic, Judaic, and Christian teachings are used by many people
to support the idea that people are given the sacred responsibility to be
guardians of nature (Figure 1.3; see Chapter 3, “Enviornmental Ethics”).
Many of the leaders of the early Western environmental movement that
helped to establish parks and wilderness areas did so because of strong
personal convictions that developed from their Christian religious beliefs.
n to Conservation Biology 1E Primack/Sher
Contemporary religious leaders have pointed out that some of the most
sociates
udio/SA profound moments in the Bible occur on mountaintops, in the wilderness,
her1E_01.02 or onDate
the10-29-10
banks of a river (Korngold 2008). In Native American tribes of
the Pacific Northwest, hunters undergo purification rituals in order to be
considered worthy, and the Iroquois consider how their actions would af-
fect the lives of their descendants after seven generations.
8 Chapter 1

Figure 1.3 Religious convictions


combined with a history of traditional
relationships with nature motivate the
grassroots conservation organization
Kakamega Environmental Education
Program (KEEP), established in 1998
to protect one of the last remnants of
tropical forest left in Kenya, East Africa.
(Photograph by Anna Sher.)

Modern conservation biology in the United States can be traced to the


nineteenth-century transcendentalist philosophers Ralph Waldo Emerson
and Henry David Thoreau, who saw wild nature as an important element
in human moral and spiritual development. Emerson (1836) saw nature
as a temple in which people could commune with the spiritual world and
achieve spiritual enlightenment. Thoreau was both an advocate for na-
ture and an opponent of materialistic society, writing about his ideas and
experiences in Walden. His book, published in 1854, has influenced many
generations of students and environmentalists. Eminent American wilder-
ness advocate John Muir used the transcendental themes of Emerson and
Thoreau in his campaigns to preserve natural areas. According to Muir’s
preservationist ethic, natural areas such as forest groves, mountaintops,
and waterfalls have spiritual value that is generally superior to the tangible
material gain obtained by their exploitation (Muir 1901).
Subsequent leaders paved the way for conservation biology as an ap-
plied academic discipline in the United States. Gifford Pinchot, the first
Defining Conservation Biology 9

head of the US Forest Service, developed a view of nature known


as the resource conservation ethic (Ebbin 2009). He defined Preservation of natural
natural resources as the commodities and qualities found in resources, ecosystem
nature, including timber, fodder, clean water, wildlife, and even management, and
beautiful landscapes (Pinchot 1947). The proper use of natural sustainable develop-
resources, according to the resource conservation ethic, is what- ment are major themes in
ever will further “the greatest good of the greatest number [of conservation biology.
people] for the longest time.” From the perspective of conserva-
tion biology, sustainable development is development that best
meets present and future human needs without damaging the
environment and biodiversity (Czech 2008).
Shortly thereafter, biologist Aldo Leopold published A Sand County
Almanac (1949), which also illustrated the interrelatedness of living things
and their environment, promoting the idea that the most important goal of
conservation is to maintain the health of natural ecosystems and ecological
processes (Leopold 2004). As a result, Leopold and many others lobbied
successfully for certain parts of national forests to be set aside as wilder-
ness areas (Shafer 2001). This idea of considering the ecosystem as a whole,
including human populations, is now termed the land ethic.
Marine biologist Rachel Carson (Figure 1.4) is credited with raising
public awareness of the complexity of nature with her best-selling books,
including Silent Spring (1962), which brought attention to the dangers of
pesticides and spurred an international environmental movement. An ap-
proach known as ecosystem management, which combines ideas of Car-
son, Leopold, and Pinchot, places the highest man-
agement priority on cooperation among businesses,
conservation organizations, government agencies,
private citizens, and other stakeholders to provide for
human needs while maintaining the health of wild
species and ecosystems.
Depictions of nature in the creative and perform-
ing arts have also played an important role in the
growing awareness of the value of nature and its pres-
ervation in the United States. In the mid-nineteenth
century, prolific painters of the Hudson River School
were noted for their romantic depictions of “scenes
of solitude from which the hand of nature has never
been lifted” (Cole 1965). Photographer Ansel Adams

FIGURE 1.4 Rachel Carson (1907–1964) was a marine


biologist who, through her popular writing, including Silent
Spring (1962), helped to found both conservation biology
and the environmental movement. (Courtesy of the Bei-
necke Rare Book and Manuscript Library, Yale University.)
10 Chapter 1

(1902–1984) took breathtaking images of wild America, helping to foster


public support for its protection (Figure 1.5). Popular singer-songwriter
John Denver also inspired interest in conservation with such lyrics as:
Now he walks in quiet solitude the forest and the streams, seeking grace
in every step he takes… His life is filled with wonder but his heart still
knows some fear of a simple thing he cannot comprehend: why they try
to tear the mountains down … more people, more scars upon the land.
Rocky Mountain High, 1975
Clearly, the arts play a unique role in fostering interest in conservation
(Jacobson et al. 2007).
In Europe, expressions of concern for the protection of wildlife began
to spread in the nineteenth century when the expansion of agriculture
and the use of firearms for hunting led to a marked reduction in wild
animals (Galbraith et al. 1998). These dramatic changes stimulated the
formation of the British conservation movement, leading to the founding
of the Commons, Open Spaces and Footpaths Preservation Society in

FIGURE 1.5 Ansel Adams (1902–1984) showed the public the beauty of wild
spaces. (Commissioned by the National Park Service, The Tetons, Snake River.)
Defining Conservation Biology 11

Table 1.1 Some Useful Units of Measurement


Length
1 meter (m) 1 m = 39.4 inches = ~3.3 feet
1 kilometer (km) 1 km = 1000 m = 0.62 mile
1 centimeter (cm) 1 cm = 1/100 m = 0.39 inches
1 millimeter (mm) 1 mm = 1/1000 m = 0.039 inches

Area
1 square meter (m2) Area encompassed by a square, each side of which
is 1 meter
1 hectare (ha) 1 ha = 10,000 m2 = 2.47 acres
100 ha = 1 square kilometer (km2)

Mass
1 kilogram (kg) 1 kg = 2.2 pounds
1 gram (g) 1 g = 1/1000 kg = 0.035 ounces
1 milligram (mg) 1 mg = 1/1000 g = 0.000035 ounces

Temperature
degree Celsius (°C) °C = 5/9(°F – 32)
0°C = 32° Fahrenheit (the freezing point of water)
100°C = 212° Fahrenheit (the boiling point of water)
20°C = 68° Fahrenheit (“room temperature”)

1865, the National Trust for Places of Historic Interest or Natural Beauty
in 1895, and the Royal Society for the Protection of Birds in 1899. Alto-
gether, these groups have preserved nearly 1 million hectares (ha) of open
land (Table 1.1 provides an explanation of the term hectare and other
measurements). In contrast to its origins in the United States, biological
conservation in Europe has had a more integrated view of human society
and ecosystems as a whole, rather than envisioning a dichotomy of man
versus nature (Linnell et al. 2015).
Many societies worldwide similarly have strong traditions
of nature conservation and land protection. Tropical countries
such as Brazil, Costa Rica, and Indonesia have a history of rever-
As demonstrated by the
ence for nature, and their governments have allocated increasing
numbers and areas of national parks. The economic value of conservation tradition
these protected areas is constantly increasing because of their in Europe, habitat
importance for tourism and the valuable ecosystem services degradation and species
they provide, such as purifying water and absorbing carbon di- loss can catalyze long-
oxide (see Chapter 3). Many tropical countries have established lasting conservation
agencies to regulate the exploration and use of their biodiversity, efforts.
and these efforts increasingly involve the indigenous peoples
12 Chapter 1

who depend on and have unique knowledge of these ecosystems. Hunt-


ing and gathering societies, such as the Penan of Borneo, give thousands
of names to individual trees, animals, and places in their surroundings
to create a cultural landscape that is vital to the well-being of the tribe.
Indeed, traditional societies throughout the world have influenced and
enriched modern conservation biology.

A new science is born


By the early 1970s, scientists throughout the world were aware of an ac-
celerating biodiversity crisis, but there was no central forum or organi-
zation to address the issue. Scientist Michael Soulé organized the first
International Conference on Conservation Biology in 1978 so that wild-
life conservationists, zoo managers, and academics could discuss their
common interests. At that meeting, Soulé proposed a new interdisciplin-
ary approach that could help save plants and animals from the threat
of human-caused extinctions, which he called conservation biology (Soulé
1985). Subsequently, Soulé, along with colleagues including Paul Ehrlich
of Stanford University and Jared Diamond of the University of California
at Los Angeles, began to develop conservation biology as a discipline that
would combine the practical experience of wildlife, forestry, fisheries,
and national park management with the theories of population biology
and biogeography. In 1985, this core of scientists founded the Society for
Conservation Biology. This organization, which continues today with an
international membership, sponsors conferences and publishes the sci-
entific journal Conservation Biology. That journal is now joined by many
peer-reviewed publications that feature research in conservation biology,
expanding our understanding of the importance of biodiversity and how
it can be protected.
Public and policymaker awareness of the value of, and threats to, biodi-
versity greatly increased following the international Earth Summit held in
Rio de Janeiro, Brazil, in 1992 (see Chapter 11). At this meeting, representa-
tives of 178 countries formulated and eventually signed the Convention on
Biological Diversity (CBD), which obligates countries to protect their biodi-
versity but also allows them to obtain a share in the profits of new prod-
ucts developed from that diversity. In 2000, the United Nations
General Assembly adopted May 22 as International Biodiversity
Day to commemorate the conference, and in 2006, the US Con-
Interdisciplinary gress designated the third Friday in May as Endangered Species
approaches, the involve- Day. The value of biodiversity was further highlighted by the
ment of local people, United Nations designation of 2010 as the International Year of
and the restoration of Biodiversity and 2011 as the International Year of Forests. In 2015,
important environments the United Nations Climate Change Conference (COP21) was at-
and species all attest to tended by 196 parties and resulted in the Paris Agreement to
progress in the science of reduce global greenhouse emissions (Figure 1.6). Arguably, this
conservation biology. increase in understanding and concern would not have been pos-
sible without the foundation of a recognized scientific discipline.
Defining Conservation Biology 13

FIGURE 1.6 The heads of delegations at the 2015 United Nations Conference
on Climate Change in Paris (also referred to as COP21). President Barak Obama
is in the center, speaking to another delegate head. (© Presidencia de la Repub-
lica Mexicana/Flickr, licensed under a Creative Commons Attribution license.)

The interdisciplinary approach: A case study with sea turtles


Throughout the world, scientists are using the approaches of conservation
biology to address challenging problems, as illustrated by the efforts to
save the Kemp’s ridley sea turtle (Lepidochelys kempii). The Kemp’s ridley is
the rarest and smallest of the world’s sea turtle species, at 70–100 cm (2–3
feet) long and about 45 kg (100 pounds). This critically endangered spe-
cies is now recovering as a result of international conservation efforts and
cooperation between scientists, conservation organizations, government
officials, and the interested public.
After its discovery as a distinct species in the late 1800s, it took nearly
a century for scientists to determine how and where the Kemp’s ridley
reproduces (Wibbels and Bevan 2015). They discovered that nearly 95%
of Kemp’s ridley nesting happens on beaches in the state of Tamaulipas,
in the northeastern corner of Mexico, in highly synchronized gatherings
of turtles, called arribada. This highly concentrated breeding is unusual
among turtles and makes the species particularly vulnerable to inten-
sive harvesting. Over many decades, locals collected an estimated 80%
of Kemp’s ridley eggs from the nesting beaches for eating. Thousands of
turtles also drowned in fishing gear, especially in shrimp nets. In 1985,
the progressive decline in the population brought it to a low point of only
14 Chapter 1

702 nests worldwide, making the Kemp’s ridley the most endangered sea
turtle in the world.
Heeding the warning of wildlife biologists that the species was nearing
extinction, government officials from Mexico and the United States worked
together to help the species recover and establish stable populations. As a
first step, nesting beaches were protected as refuges, reserves, and parks.
Egg collection was banned. And at sea, shrimp trawlers were required
to use turtle excluder devices (TEDs), consisting of a grid of bars with an
opening that allows a caught turtle to escape.
In addition to reducing threats, a collaborative group of national and
state agencies and conservation organizations in Mexico and the United
States has undertaken an ambitious effort to increase nest and hatch-
ling survival and to improve education and appreciation of sea turtle
conservation. In the United States, national park
authorities began to reestablish a population on
Padre Island in Texas, where the species had
formerly occurred. From 1978 to 1988, scientists,
conservationists, and volunteers collected 22,507
eggs from Mexico, packed them in sand, and trans-
ported them to Padre Island National Seashore,
which is managed by the US National Park Service
(Figure 1.7). The hatchlings were released on the
beach and briefly allowed to swim in the surf be-
fore they were captured using aquarium dip nets.
The hope was that this brief time on the beach and
in the surf would help them imprint on the site
so that they would return there to nest as adults.
The captured hatchlings were then reared in cap-
tivity for 9–11 months as a part of a “head-start”
program that allowed the turtles to grow large
enough to avoid most predators. (Most sea turtles
die as hatchlings.) Scientists carefully monitored
growth during this period (Caillouet et al. 1997).
Then the one-year-old turtles were released per-
manently into the Gulf of Mexico (see the chapter
opening photo).
Now, each year, the staff at Padre Island, many
partner organizations, and over a hundred volun-
teers patrol the beach during the breeding season,
searching for Kemp’s ridleys and their nests. When
Figure 1.7 Researchers collect eggs from they find nests, teams carefully excavate them and
a Kemp’s ridley sea turtle nest. The eggs will bring the eggs to an incubation facility or a large
either be relocated to a nest within a protected screen enclosure called a corral. When the young
enclosure or brought to an incubation facility. hatchlings are released, it is now a public event
(Courtesy of David Bowman, US Fish and Wild- that doubles as an education tool—the hope is that
life Service.) the people watching each release will become ad-
Defining Conservation Biology 15

vocates for the turtles’ protection. Outside the national seashore, private
conservation organizations also help protect the turtles on their feeding
grounds. Together, these conservation activities and associated media cov-
erage expose hundreds of thousands of visitors to information about sea
turtle ecology and conservation.
Over a 16-year period, the Kemp’s ridley population at Padre Island
National Seashore increased dramatically, from 6 nests, 590 eggs, and 369
hatchlings released in 1996 to 209 nests, 20,067 eggs, and 16,577 hatchlings
released in 2012. Compared with the low of 702 nests in 1985, researchers
and volunteers counted a total of 21,797 nests in 2012. Each female lays two
to three clutches of eggs each season, so this number of nests corresponds
to at least 7000–9000 mature reproducing females. However, the number
declined to only 12,053 nests in 2014, at least partially due to the Deepwater
Horizon oil spill in 2010 that is believed to have killed hundreds of juvenile
turtles (Caillouet et al. 2015). Fortunately, a study of the nests found preda-
tion rates to be low and hatchling survival high (Bevan et al. 2014), and
the number of nests grew by more than a thousand in the next year (Luis
Jaime Peña, pers. comm.).
The Kemp’s Ridley Sea Turtle Recovery Plan has set a target of 10,000
nesting females for the population to be considered recovered. Scientific
scrutiny, international partnerships, and the participation of volunteers
and local communities have brought the Kemp’s ridley sea turtle back from
the brink of extinction, and all of those involved will continue to seek the
answers leading to its complete recovery.

The Ethical Principles of


Conservation Biology
Earlier in the chapter, we mentioned that conservation biology is a norma-
tive discipline in which certain value judgments are inherent. The field
rests on an underlying set of principles that is generally agreed on by
practitioners of the discipline (Soulé 1985; “Organizational Values,” sensu
Society for Conservation Biology 2016) and can be summarized as follows:
1. Biological diversity has intrinsic value. Species and the biological commu-
nities in which they live possess value of their own, regardless of their
economic, scientific, or aesthetic value to human society. This
value is conferred not just by their evolutionary history and
unique ecological role, but also by their very existence. (See
Chapter 3 for a more complete discussion of this topic.) There are ethical reasons
2. The untimely extinction of populations and species should be pre- why people want to
vented. The ordinary extinction of species and populations conserve biodiversity,
as a result of natural processes is an ethically neutral event. such as belief that
In the past, the local loss of a population was usually off- species have intrinsic
set by the establishment of a new population through dis- value.
persal. However, as a result of human activity, the loss of
16 Chapter 1

populations and the extinction of species has increased by more than


a hundredfold, with no simultaneous increase in the generation of new
populations and species (MEA 2005; see Chapter 5).
3. The diversity of species and the complexity of biological communities should
be preserved. In general, most people agree with this principle simply
because they appreciate biodiversity; it has even been suggested that
humans may have a genetic predisposition to love biodiversity, called
biophilia (Figure 1.8) (Corral-Verdugo et al. 2009). Many of the most
valuable properties of biodiversity are expressed only in natural envi-
ronments. Although the biodiversity of species may be partially pre-
served in zoos and botanical gardens, the ecological complexity that
exists in natural communities will be lost without the preservation
of natural areas (see Chapter 8). Furthermore, biodiversity has been
directly linked to ecosystem productivity and stability (Hautier et al.
2015), among other values (see Chapter 3).
4. Science plays a critical role in our understanding of ecosystems. It is not
enough to simply value diversity and protect natural spaces; objective
research is necessary to identify which species and environments are
at greatest risk, as well as to understand the nature of these risks and
how to mitigate them. Ideally, scientists are involved in all stages of

FIGURE 1.8 People enjoy seeing the diversity of life, as illustrated by the popularity
of planting gardens and of public botanical gardens as tourist destinations. (Butchart
Gardens, Victoria, BC, Canada © Xuanlu Wang/Shutterstock.)
Defining Conservation Biology 17

conservation, including implementation of conservation actions and


monitoring results.
5. Collaboration among scientists, managers, and policy makers is necessary.
In order to achieve the goals of reduced extinction rates and preserva-
tion of biological communities, high-quality research findings must
be shared with those who create the laws and provide the funding for
conservation actions as well as those who must implement them. These
actions are most likely to succeed when they are based on scientifically
sound information.

Not every conservation biologist accepts every one of these principles,


and there is no hard-and-fast requirement to do so. Individuals or organiza-
tions that agree with even two or three of these principles are often willing
to support conservation efforts. Current progress in protecting species
and ecosystems has been achieved in part through partnerships between
traditional conservation organizations such as The Nature Conservancy
and cattle ranchers, hunting clubs like Ducks Unlimited, and other groups
with a vested interest in the health of ecosystems.

Looking to the Future


The field of conservation biology has set itself some imposing—and abso-
lutely critical—tasks: to describe Earth’s biological diversity, to protect what
remains, and to restore what is degraded. The field is growing in strength,
as indicated by increased governmental participation in conservation ac-
tivities, increased funding of conservation organizations and projects, and
an expanding professional society.
In many ways, conservation biology is a crisis discipline. Decisions
about selecting national parks, species management, and other aspects of
conservation are made every day under severe time pressure (Laurance et
al. 2012; Martin et al. 2012). As one of the guiding values mentioned above,
biologists and scientists in related fields seek to provide the advice that
governments, businesses, and the general public need in order to make
crucial decisions, but because of time constraints, scientists are often com-
pelled to make recommendations without thorough investigation. Deci-
sions must be made, with or without scientific input, and conservation
biologists must be willing to express opinions and take action based on the
best available evidence and informed judgment (Maron et al. 2013). They
must also articulate a long-term conservation vision that extends beyond
the immediate crisis (Wilhere 2012).
Despite the threats to biodiversity and the limitations of our knowledge,
we can detect many positive signs that allow conservation biologists to
be cautiously hopeful (Roman et al. 2015). The rate of human population
growth has slowed (US Census Bureau), and per capita energy use in the
United States, while still high, has been decreasing; in 2014 it was the lowest
it has been since the 1960s (World Bank). The number of protected areas
18 Chapter 1

around the globe continues to increase, with a dramatic expansion in the


number of marine protected areas and a commitment by coastal nations
to increase them fivefold by 2020 (Halpern 2014).
These gains are due in part to action spurred by the public. Increasing
numbers of social and religious leaders have rallied for the protection of
biodiversity. As one powerful example, the first papal encyclical focused
solely on the environment was recently released, in which our obligation to
“till and keep” the garden of the world (cf. Gen. 2:15) was explained thus:
“Tilling” refers to cultivating, … while “keeping” means caring,
protecting, overseeing and preserving. This implies a relationship
of mutual responsibility between human beings and nature. Among
positive experiences in this regard, we might mention, for example, …
the binding Convention on International Trade in Endangered Species
of Wild Fauna and Flora, which includes on-site visits for verifying
effective compliance.
Pope Francis, Laudato Si: On Care for Our Common Home, 2015
Our ability to protect biodiversity has been strengthened by a wide
range of local, national, and international efforts. Many endangered species
are now recovering as a result of such conservation measures (IUCN 2015).
Effective action has resulted from our continuing expansion of knowledge
in conservation science, the developing linkages with rural development
and social sciences, and our increased ability to restore degraded environ-
ments. All of these advances suggest that progress is being made, despite
the enormous tasks still ahead.

Summary
„„Human activities are causing the extinc- „„Elements of conservation biology can
tion of thousands of species both locally be found in many cultures, religions,
and globally, with threats to species and and forms of creative expression. The
ecosystems accelerating due to human modern field of conservation biology
population growth and the associated grew from the ideas of several influen-
demands for resources. tial individuals, eventually becoming a
„„Conservation biology is a field that com- recognized scientific discipline with pro-
bines basic and applied disciplines with fessional societies and academic journals
three goals: to describe the full range by the 1980s.
of biodiversity on Earth, to understand „„Conservation biology rests on a num-
human impact on biodiversity, and to ber of underlying assumptions that are
develop practical approaches for pre- accepted by most professionals in the
venting species extinctions, maintaining discipline: biodiversity has value in and
genetic diversity, and protecting and of itself, extinction from human causes
restoring ecosystems. should be prevented, diversity at mul-
Defining Conservation Biology 19

tiple levels should be preserved, science are many successful projects, such as the
plays a critical role, and scientists must conservation of Kemp’s ridley sea turtles,
collaborate with nonscientists to achieve that indicate that progress can be made.
our goals.
„„The conservation of biodiversity has be-
come an international undertaking. There

For Discussion
1. How is conservation biology funda- with which you are familiar. Do you
mentally different from other branches think their guiding philosophies are clos-
of biology, such as physiology, genetics, est to the resource conservation ethic,
or cell biology? How is it similar to the the preservation ethic, or the land ethic?
science of medicine? How is it different What factors allow them to be successful
from environmentalism? or limit their effectiveness? Learn more
2. What do you think are the major conser- about these organizations through their
vation and environmental problems fac- publications and websites.
ing the world today? What are the major 4. How would you characterize your own
problems facing your local community? viewpoint about the conservation of bio-
What ideas for solving these problems diversity and the environment? Which of
can you suggest? (Try answering this the religious or philosophical viewpoints
question now, and once again when you of conservation biology stated here do
have completed this book.) you agree or disagree with? How do you,
3. Consider the public land management or could you, put your viewpoint into
and private conservation organizations practice?

Suggested Readings
Barnosky, A. D. and 11 others. 2011. Has Earth’s sixth mass extinction already
arrived? Nature 471: 51–57. Evidence from the fossil records and modern
extinction rates suggest that we are on the verge of a major extinction
event.
Bevan, E. and 11 others. 2014. In situ nest and hatchling survival at Rancho
Nuevo, the primary nesting beach of the Kemp’s ridley sea turtle, Lepido-
chelys kempii. Herpetological Conservation and Biology 9: 563–577. Research
shows that this endangered species has a promising future.
Bradshaw, C. J. A. and B. W. Brook. 2014. Human population reduction is not a
quick fix for environmental problems. Proceedings of the National Academy of
Sciences 111: 16610–16615.
Caillouet, C. W., B. J. Gallaway, and A. M. Landry. 2015. Cause and call for
modification of the bi-national recovery plan for the Kemp’s ridley sea
turtle (Lepidochelys kempii)—Second Revision. Marine Turtle Newsletter 145:
1–4.
Carson, R. 1962. Silent Spring. Houghton Mifflin Company, Boston. Essays
written over a period from 1958–1962 on the devastating effects of pesti-
cide use on ecosystems, particularly birds.
20 Chapter 1

Halpern, B. S. 2014. Making marine protected areas work. Nature 506: 167–168.
Hautier, Y., D. Tilman, F. Isbell, E. W. Seabloom, E. T. Borer, and P. B. Reich.
2015. Anthropogenic environmental changes affect ecosystem stability via
biodiversity. Science 348: 336–340. Of several factors explored, it was only
those that decreased biodiversity that affected the stability of ecosystem
productivity.
Hitzhusen, G. E. and M. E. Tucker. 2013. The potential of religion for Earth
Stewardship. Frontiers in Ecology and the Environment 11: 368–376. There is
a natural alliance between religion and conservation biology that can be a
positive force.
Horton, C., T. R. Peterson, P. Banerjee, and M. J. Peterson. 2015. Credibility
and advocacy in conservation science. Conservation Biology doi: 10.1111/
cobi.12558. The nature of conservation biology as a normative science
means that there can be a potential conflict between the dispassionate na-
ture of science and the interest in promoting conservation policy.
Kloor, K. 2015. The battle for the soul of Conservation Science. Issues in Science
and Technology 31: 74. The ongoing debate between those who believe that
the field should be primarily guided by “nature for nature’s sake” and oth-
ers who believe that human needs should have greater weight.
Leopold, A. 1949. A Sand County Almanac. Oxford University Press, New York.
Leopold’s evocative essays articulate his “land ethic,” defining human
duty to conserve the land and the living things that thrive upon it.
Pimm, S. L. and 8 others. 2014. The biodiversity of species and their rates of
extinction, distribution, and protection. Science 344: 1246752. Most species
on the planet are not yet known to science, having restricted geographic
ranges and therefore being at risk of going extinct before they are ever
known.
Pooley, S. P., J. Andrew-Mendelsohn, and E. J. Milner-Gulland. 2014. Hunt-
ing down the chimera of multiple disciplinarity in conservation science.
Conservation Biology 28: 22–32. Even as conservation biologists attempt to
include a wider variety of disciplines, there are methodological and con-
ceptual challenges to such broad approaches.
Roman, J., M. M. Dunphy-Daly, D. W. Johnston, and A. J. Read. 2015. Lifting
baselines to address the consequences of conservation success. Trends in
Ecology and Evolution 30.6: 299–302.
Sodhi, S. N., R. Butler, W. F. Laurance, and L. Gibson. 2011. Conservation suc-
cesses at micro-, meso- and macroscales. Trends in Ecology and Evolution 26:
585–594. There are many examples of successful conservation that can be
used to guide future actions.
Soulé, M. E. 1985. “What is conservation biology?” BioScience 35: 727–734. Key
early paper defining the field, and still relevant today for its emphasis on
the intrinsic value of biodiversity.
Wibbels, T. and E. Bevan. 2015. New Riddle in the Kemp’s Ridley Saga. In State
of the World’s Sea Turtles Report. Oceanic Society.

KEY JOURNALS IN THE FIELD Biodiversity and Conservation, Biological


Conservation, BioScience, Conservation Biology, Conservation Letters, Ecological
Applications, National Geographic, Trends in Ecology and Evolution
2 What Is Biodiversity?
Species Diversity 25
Genetic Diversity 31
Ecosystem Diversity 33
Biodiversity Worldwide 41

Coral reefs are built up from the skeletons of billions of tiny individual
animals. The intricate coral landscapes create a habitat for a variety
of other marine species, including hundreds of different fish and
invertebrate creatures.
T
he protection of biological diversity is central
to conservation biology. Conservation biolo-
gists use the term biological diversity, or sim-
ply biodiversity, to mean the complete range of species
and biological communities on Earth, as well as the
genetic variation within those species and all ecosys-
tem processes. By this definition, biodiversity must be
considered on at least three levels (Figure 2.1):
1. Species diversity: All the species on Earth, including sin-
gle-celled bacteria and protists as well as the species of
the multicellular kingdoms (plants, fungi, and animals)
2. Genetic diversity: The genetic variation within species,
both among geographically separate populations and
among individuals within single populations
3. Ecosystem diversity: The different biological communi-
ties and their associations with the chemical and physical
environment (the ecosystem)
All three levels of biodiversity are necessary for the
continued survival of life as we know it, and all are im-
portant to people (Levin 2001; MEA 2005). All of these
levels are also currently facing significant threats, to
be discussed in Chapters 4 and 5, although threats to
species diversity tend to receive the most attention.
24 Chapter 2

Genetic diversity Species diversity Ecosystem diversity

FIGURE 2.1 Biological diversity includes genetic diversity (the genetic varia-
tion found within each species), species diversity (the range of species in a given
ecosystem), and ecosystem diversity (the variety of habitat types and ecosystem
processes extending over a given region).

Species diversity reflects the entire range of evolutionary and ecological


adaptations of species to particular environments. It provides people with
resources and resource alternatives; for example, a tropical rain forest or
a temperate swamp with many species produces a wide variety of plant
and animal products that can be used as food, shelter, and medicine. Ge-
netic diversity is necessary for any species to maintain reproductive vital-
ity, resistance to disease, and the ability to adapt to changing conditions
Introduction to Conservation Biology 1E Primack/Sher
Sinauer Associates
(Laikre et al. 2010). Genetic diversity is of particular value in the breeding
Morales Studio programs necessary to sustain and improve modern domesticated plants
Primack/Sher_02.01 Date 11-17-15 and animals and their disease resistance. Ecosystem diversity results from
the collective response of species to different environmental conditions.
What Is Biodiversity? 25

Biological communities found in deserts, grasslands, wetlands, and forests


support the continuity of proper ecosystem functioning, which provides
crucial services to people, such as water for drinking and agriculture, flood
control, protection from soil erosion, and filtering of air and water. We will
now examine each level of biodiversity in turn.

Species Diversity
Recognizing and classifying species is one of the major goals of conserva-
tion biology. Identifying the process whereby one species evolves into one
or more new species is one of the ongoing accomplishments of modern
biology. The origin of new species is normally a slow process, taking place
over hundreds, if not thousands, of generations. The evolution of higher
taxa, such as new genera and families, is an even slower process, typi-
cally lasting hundreds of thousands or even millions of years. In contrast,
human activities are destroying the unique species built up by these slow
natural processes in only a few decades.

What is a species?
Although seemingly a straightforward concept, how to distinguish a se-
lection of organisms as a species is subject to great scientific discussion,
and at least seven different ways of doing this have been proposed (Wiens
2007). The three most commonly used in conservation biology are:
1. Morphological species: A group of individuals that appear different from
others, that is, that are morphologically distinct. A group that is distin-
guished exclusively by such visible traits as form or structure may be
referred to as a morphospecies.
2. Biological species: A group of individuals that can potentially breed
among themselves in the wild and that do not breed with individuals
of other groups.
3. Evolutionary species: A group of individuals that share unique similari-
ties in their DNA and hence their evolutionary past.

Because the methods and assumptions used are different,


these three approaches to distinguishing species sometimes do
not give the same results. Increasingly, DNA sequences and
other molecular markers are being used to identify and distin- Using morphological
guish species that look almost identical, such as types of bacteria and genetic information
(Papadopoulou 2015). to identify species is
The morphological definition of species is the one most a major activity for
commonly used by taxonomists, biologists who specialize in taxonomists; accurate
the identification of unknown specimens and the classification identification of a species
of species (Figure 2.2). The biological definition of species is is a necessary first step
widely accepted, but it is problematic for groups of organisms in its conservation.
in which different species readily interbreed, or hybridize,
26 Chapter 2

(A) (B)

FIGURE 2.2 (A) A plant ecologist prepares a museum


specimen using a plant press. The flattened and dried
plant will later be mounted on heavy paper with a label
giving detailed collection information. (B) An ornitholo-
gist at the Museum of Comparative Zoology, Harvard
University, classifying collections of orioles: black-cowled
orioles (Icterus prosthemelas), from Mexico, and Bal-
timore orioles (Icterus galbula), which occur through-
out eastern North America. (A, photograph by Richard such as plants. Furthermore, the biologi-
Primack; B, photograph courtesy of Jeremiah Trimble, cal definition of species is difficult to use
Museum of Comparative Zoology, Harvard University; because it requires a knowledge of which
© President and Fellows of Harvard College.)
individuals actually have the potential
to breed with one another. Similarly, the
evolutionary definition of species requires
access to expensive laboratory equipment and so cannot be used in the
field. As a result, field biologists must rely on observable attributes, and
they may name a group of organisms as a morphospecies until taxonomists
can investigate them more carefully to determine if they are a distinct
species (Chan et al. 2015).
Ideally, specimens collected in the field are catalogued and stored in
one of the world’s 6500 natural history museums (for animals and other
organisms) or its 300 or more major herbaria (for plants and fungi). These
permanent collections form the basis of species descriptions and systems
of classification. Each species is given a unique two-part name (a binomial),
such as Canis lupus for the gray wolf. The first part of the name, Canis,
identifies the genus (the canids, or dogs). The second part of the name,
lupus, identifies the smaller group within the genus, the species that is
the gray wolf. This naming system both separates the gray wolf from and
connects it to similar species—such as Canis latrans, the coyote, and Canis
rufus, the red wolf.

Introduction to Conservation Biology 1E Primack/Sher


Sinauer Associates
Morales Studio/SA
What Is Biodiversity? 27

FIGURE 2.3 Cope’s gray tree frog (Hyla chrysocelis, left) is only distinguish-
able from the gray tree frog (H. versicolor, right) by their calls, but they fit both
the biological and evolutionary definition of species because they have different
numbers of chromosomes (H. versicolor is tetraploid; H. chrysocelis is diploid)
and thus are incapable of interbreeding. (Left photograph, © Jack Glisson/Alamy
Stock Photograph; right photograph, David McIntyre.)

Problems in distinguishing and identifying species are more common


than many people realize (Frankham et al. 2012). For example, a single
species may have several varieties that have observable morphological
differences, yet are similar enough to be a single biological or evolutionary
species. Different varieties of dogs, such as German shepherds, collies,
and beagles, all belong to one species; their genetic differences are actu-
ally very small, and they readily interbreed. Alternatively, closely related
“sibling” species appear very similar in morphology and physiology, yet
are genetically quite distinct (Figure 2.3).
To further complicate matters, individuals of related but distinct species
may occasionally mate and produce hybrids, intermediate forms that blur
the distinction between species. Hybridization is particularly common
among plant species in disturbed habitats. Hybridization in both plants
and animals frequently occurs when a few individuals of a rare species
are surrounded by large numbers of a closely related species. For example,
the endangered California tiger salamander (Ambystoma californiense) and
the introduced barred tiger salamander (A. mavortium) are thought to have
evolved from a common ancestor five million years ago, yet they readily
mate in California (Figure 2.4). The hybrid salamanders have a higher
fitness and are better able to tolerate environmental pollution than the
native species, A. californiense, further complicating the conservation of
this endangered species (Ryan et al. 2013).
The inability to clearly distinguish one species from another, whether
due to similarities of characteristics or to confusion over the correct scien-
tific name, often slows down efforts at species protection. It is difficult to
Introduction to Conservation Biology 1E Primack/Sher
Sinauer Associates
Morales Studio/SA
Primack_Sher1E_02.03 Date 11-09-15 03-21-16 03-22-16
28 Chapter 2

FIGURE 2.4 The hybrid ti-


ger salamander (left) is larger
than its endangered parent
species, the California tiger
salamander (right), and is in-
creasing in abundance. Note
the much larger head of the
hybrid salamander. (Photo-
graph courtesy of H. Bradley
Shaffer.)

write precise, effective laws to protect a species if scientists and lawmakers


are not certain which individuals belong to which species. At the same
time, species are going extinct before they are even described. Tens of
thousands of new species are being described each year, but even this rate
is not fast enough. The key to solving this problem is to train more tax-
onomists, especially for work in the species-rich tropics (Joppa et al. 2011).
Those conservation biologists primarily concerned with ecosystem
function rather than individual species extinction have argued that a better
measure than species diversity is functional diversity; that is, the diversity
of organisms categorized by their ecological roles or traits rather than their
taxonomy (Gagic et al. 2015). Functional diversity is an especially important
concept in the context of habitat restoration (see Chapter 10). However, if
our goal is to prevent untimely extinctions, we cannot avoid the task of
identifying and measuring species diversity.

Measuring species diversity


Conservation biologists often want to identify locations of high species di-
versity. Quantitative definitions of species diversity have been developed
by ecologists as a means of comparing the overall diversity of different
communities at varying geographic scales (Flohre et al. 2011).
At its simplest level, species diversity can be defined as the number of
species, called species richness. This number can be determined by several
methods and at different geographic scales. Three diversity measurements
are based on species richness:
• Alpha diversity is the number of species found in a given community,
such as a lake or a meadow.
• Gamma diversity is the number of species at larger geographic scales
that include a number of ecosystems, such as a mountain range or a
continent.
What Is Biodiversity? 29

• Beta diversity links alpha and gamma diversity and represents the
rate of change of species composition as one moves across a large region. For
example, if every lake in a region contained a similar array of fish
species, then beta diversity would be low; on the other hand, if the
bird species found in one forest were entirely different from the bird
species in separate but nearby forests, then beta diversity would be
high. There are several ways of calculating beta diversity; a simple
measure of beta diversity can be obtained by dividing gamma diver-
sity by alpha diversity.
We can illustrate these three types of diversity with a theoretical ex-
ample of three mountain ranges (Figure 2.5). Region 1 has the highest
alpha diversity, with more species per mountain on average (six species)

Region 1 ALPHA GAMMA BETA


(species per (species per (gamma/
mountain) region) alpha)
DE DE DE
C F C F C F 6 7 1.2
B B B G
A A

Region 2

BC EF HI
4 10 2.5
A D G G J
D

Region 3

B
A D E FG
C E H 3 8 2.7

FIGURE 2.5 Biodiversity indexes for three regions, each consisting of three
separate mountains. Each letter represents a population of a species; some spe-
cies are found on only one mountain, while other species are found on two or
three mountains. Alpha, gamma, and beta diversity values are shown for each
region. If funds were available to protect only one region, Region 2 should be
selected because it has the greatest gamma (total) diversity. However, if only
one mountain could be protected, a mountain in Region 1 should be selected
because these mountains have the highest alpha (local) diversity, that is, the
greatest average number of species per mountain. Each mountain in Region 3
has a more distinct assemblage of species than the mountains in the other two
regions, as shown by the higher beta diversity. If Region 3 were selected for pro-
tection, the relative priority of the individual mountains should then be judged
based on how many unique species are found on each mountain.
30 Chapter 2

Figure 2.6 If each circle represents a random sample of fish from a pond
and colors represent species, both have the same species richness: 5. However,
pond sample A is dominated by a single species (6 orange fish out of a total of 10
fish or 60% of the total), while each of the other four species has only 10% of the
total. In contrast, pond sample B has perfect evenness, that is, each of the five
species has the same number of individuals, or 20% of the total. Therefore we
would consider pond B to have greater species diversity. We can further quantify
this by calculating H, a measure of diversity, as shown in each table, with pond A
having a diversity of 0.53 and pond B having a diversity of 0.70.

than the other two regions. Region 2 has the highest gamma
Identifying patterns of diversity, with a total of 10 species. Dividing gamma diversity by
species diversity helps alpha diversity shows that Region 3 has a higher beta diversity
conservation biologists (2.7) than Region 2 (2.5) or Region 1 (1.2) because all of its species
establish which locations are found on only one mountain each. In practice, indexes of
are most in need of diversity are often highly correlated. The plant communities of
the eastern foothills of the Andes, for instance, show high levels
protection.
of diversity at alpha, beta, and gamma scales.
More complex indexes, such as the Shannon diversity index
(also called the Shannon-Wiener index), Simpson index, and
Pielou evenness index, take the relative abundance of differ-
ent species into account; by these measures, a community dominated by
a few species is less diverse than one with a more even distributions of
species, even with the same species richness. The Shannon diversity index
is calculated as
H = – ∑[ p i × ln ( pi ) ]
that is, the negative sum of the proportion (p) of each species (i) multiplied
by the natural log (ln) of p. In a simple example, let’s image two ponds,
each of which has five fish species. In pond A, 60% of the individuals
are orange carp and each of the remaining four species only represent
10% of the individuals, whereas in pond B there are also five fish species
but each of them has equal numbers of individuals, or 20% of the total.
Using the Shannon diversity index, pond B will have a greater diversity
than pond A (Figure 2.6). In some cases, one pond may even have a
greater number of species but a lower diversity index than another pond
if its community is dominated by one or a few particular species. Note
that, like the richness values explained above, diversity measures of this
type can be calculated at different scales and therefore are useful only
as relative, rather than absolute, values. Furthermore, these quantitative
definitions of diversity capture only part of the broad definition of bio-
diversity used by conservation biologists, and new ones continue to be
developed (Iknayan et al. 2014; Magurran and McGill 2013). Although
each has its limitations, they are useful for comparing regions and high-
lighting areas that have large numbers of native species requiring con-
servation protection.
What Is Biodiversity? 31

(A) (B)

Number n/ total Number n/ total


in = % in in = % in
sample sample Log Log (p) sample sample Log Log (p)
(n) (p) (p) × (p) (n) (p) (p) × (p)

6 0.60 –0.22 –0.13 2 0.20 –0.70 –0.14

1 0.10 –1.00 –0.10 2 0.20 –0.70 –0.14

1 0.10 –1.00 –0.10 2 0.20 –0.70 –0.14

1 0.10 –1.00 –0.10 2 0.20 –0.70 –0.14

1 0.10 –1.00 –0.10 2 0.20 –0.70 –0.14

Total 10 1 –0.53 Total 10 1 –0.70


fish (100%) fish (100%)
H = 0.53 H = 0.70

Genetic Diversity
Conservation biology also concerns itself with the preservation of genetic
diversity within a species. This level of diversity is important because it
provides evolutionary flexibility: when environmental conditions change,
a genetically diverse species is more likely to have traits that allow it to
adapt. Rare species often have less genetic variation than widespread
species and, consequently, are more vulnerable to extinction (Frankham
et al. 2009; see Chapter 5).
Genetic diversity is important both within and among populations.
A population is a group of individuals that mate with one another and
produce offspring; species may contain one or many populations.
Genetic diversity arises because individuals have slightly different
forms of their genes (or loci), the units of the chromosomes that code for

Introduction to Conservation Biology 1E Primack/Sher


32 Chapter 2

Two different Within-population variation


alleles of gene A
Among-population variation
Individual

A a FIGURE 2.7 There is genetic variation


Chromosomes
within individuals due to variation in the
B B
(one from each parent) alleles of particular genes and variation be-
Individual genes tween chromosomes. There is also genetic
A and B variation between individuals within popu-
lations and among separate populations.
Within-individual variation
(After Groom et al. 2006.)

specific proteins. These different forms of a gene are known as alleles, and
the differences originally arise through mutations—changes that occur in
the deoxyribonucleic acid (DNA) that constitutes an individual’s chromo-
somes. Genetic variation increases when offspring receive unique combina-
tions of genes and chromosomes from their parents via the recombination
of genes that occurs during sexual reproduction. Genes are exchanged
between chromosomes, and new combinations are created when chromo-
somes from two parents combine to form a genetically unique offspring.
Although mutations provide the basic material for genetic variation, the
random rearrangement of alleles in different combinations that character-
izes sexually reproducing species dramatically increases the potential for
genetic variation (Figure 2.7).
The total array of genes and alleles in a population is the gene pool
of the population, while the particular combination of alleles that any
individual possesses is its genotype (Winker 2009). The phenotype of an
individual represents the morphological, physiological, anatomical, and
biochemical characteristics of that individual that result from the expres-
sion of its genotype in a particular environment. Examples of phenotypes
include eye color and blood type, physical qualities that are determined
predominantly by an individual’s genotype.
The amount of genetic variation in a population is determined by both
the number of genes that have more than one allele (polymorphic genes)
and
Introduction to Conservation Biology 1E the number of alleles for each of these genes. The existence of a poly-
Primack/Sher
Sinauer Associates morphic gene also means that some individuals in the population will
Morales Studio/SA be heterozygous for the gene; that is, they will receive a different allele of
Primack_Sher1E_02.07 Date 11-09-15
What Is Biodiversity? 33

the gene from each parent. On the other hand, some individu-
als will be homozygous: they will receive the same allele from Genetic variation within
each parent. All these levels of genetic variation contribute to a species can allow
a population’s (and therefore a species’) ability to adapt to a the species to adapt to
changing environment. environmental change;
genetic variation can
Ecosystem Diversity also increase the value of
domesticated species to
Ecosystems are diverse, and this diversity is apparent even
people.
across a particular landscape. As we climb a mountain, for ex-
ample, the structure of the vegetation and the kinds of plants
and animals present gradually change from those found in a tall
forest to those found in a low, moss-filled forest to alpine meadow to cold,
barren rock. As we move across the landscape, physical conditions (soil,
temperature, precipitation, and so forth) change. One by one, the species
present at our starting point drop out, and we encounter new species that
were not found there. The landscape as a whole is dynamic and changes
in response to the overall environment and the types of human activities
that are associated with it.

What are communities and ecosystems?


A biological community is defined as the species that occupy a particular
locality and the interactions among those species. A biological community,
together with its associated physical and chemical environment, is termed
an ecosystem (Figure 2.8). Many characteristics of an ecosystem result
from ongoing processes, including water cycles, nutrient cycles, and en-
ergy capture. These processes occur at geographic scales that range from
square meters to hectares and all the way to regional scales involving tens
of thousands of square kilometers (see Table 1.1 for definitions of these met-
ric terms). For example, in a temperate forest, rain falls and is absorbed by
the soil. Some of that rain evaporates from the surface, some percolates to
groundwater reserves, and some is taken up by plants that use it in pho-
tosynthesis, converting atmospheric CO2 into carbohydrates. These plants
may then be eaten by animals, which convert the carbohydrates back to
energy through respiration, which releases CO2 back into the environment.
Other plants may decompose on the forest floor, releasing nutrients and
providing energy for bacteria, fungi, and animals.
The physical environment, especially annual cycles of temperature and
precipitation and the characteristics of the land surface, affects the struc-
ture and characteristics of a biological community and profoundly influ-
ences whether a site will support a forest, grassland, desert, or wetland.
In aquatic ecosystems, physical characteristics such as water turbulence
and clarity, as well as water chemistry, temperature, and depth, affect
the characteristics of the associated biota (a region’s flora and fauna). In
turn, the biological community can alter the physical characteristics of an
environment. For example, wind speeds are lower and humidity is higher
34 Chapter 2

Abiotic environment
Energy lost as heat Solar radiation,
water, oxygen, carbon
Not absorbed dioxide, minerals
by producers

Primary producers
Photosynthetic
species

Primary consumers
Herbivores obtain
energy from photo-
synthetic species

Secondary
consumers
Predators and
parasites feed
on herbivores

Decomposers
(Detritivores)
Scavengers feed on dead
tissues and wastes

FIGURE 2.8 A model of a field ecosystem, showing its trophic levels and sim-
plified energy pathways.

Introduction to Conservation Biology 1E Primack/Sher


What Is Biodiversity? 35

inside a forest than in a nearby grassland. Marine communi-


ties such as kelp forests and coral reefs can affect the physical Within a community,
environment as well by buffering wave action. each species has its own
Within a biological community, species play different roles requirements for food,
and differ in what they require to survive (Roscher et al. 2015). temperature, water, and
For example, a given plant species might grow best in one type other resources, any
of soil under certain conditions of sunlight and moisture, be of which may limit its
pollinated only by certain types of insects, and have its seeds population size and its
dispersed by certain bird species. Similarly, animal species dif- distribution.
fer in their requirements, such as the types of food they eat
and the types of resting and breeding places they prefer, col-
lectively referred to as habitat. Even though a forest may be
full of vigorously growing green plants, an insect species that feeds on
only one rare and declining plant species may be unable to develop and
reproduce because it cannot get the specific food that it requires. Any of
these requirements may become a limiting resource when it restricts the
size of a population.

Species interactions within ecosystems


The composition of ecosystems is often affected by competition and pre-
dation (Cain et al. 2014). Predators are animals that hunt and eat prey,
which are the organisms that are eaten. Predation on plants is generally
referred to as herbivory. Predators of all types may dramatically reduce
the densities of certain prey species and even eliminate some species from
particular habitats. Indeed, predators may indirectly increase the number
of prey species in an ecosystem by keeping the density of each species so
low that severe competition for resources does not occur.
In many ecosystems, predators keep the number of individuals of a
particular prey species below the number that the resources of the ecosys-
tem can support, a number termed the habitat’s carrying capacity. If the
predators (e.g., wolves; Canis lupus) are removed by hunting, poisoning,
or some other human activity, the prey population (e.g., deer; Odocoileus
spp.) may increase to carrying capacity, or it may increase beyond carry-
ing capacity to a point at which crucial resources are overtaxed and the
population crashes. In addition, the population size of a species may be
controlled by other species that compete with it for the same resources;
for example, the population size of terns that nest on a small island may
decline or grow if a gull species that uses the same nesting sites becomes
abundant or is eliminated from the site.
Community composition is also affected when two species benefit each
other in a mutualism (Figure 2.9). Mutualistic species reach higher densi-
ties when they occur together than when only one of the species is present.
One example of mutualism is the relationship between fruit-eating birds
and plants with fleshy fruit containing seeds that are dispersed by birds.
Another example is flower-pollinating insects and flowering plants. In
some cases these relationships are symbiotic, and the species apparently
36 Chapter 2

FIGURE 2.9 This bohe-


mian waxwing (Bombycilla
garrulus) will disperse the fruit
of a mountain ash tree (Sorbus
aucuparia) by ingesting the
fruit and then defecating the
seed in some other location
far from the parent tree. In
this way, both species benefit.
(© All Canada Photos/Alamy
Stock Photo.)

cannot survive without each other. For example, certain symbiotic algae
living inside coral animals are ejected following unusually high water
temperatures in tropical areas, leading to the weakening and subsequent
death of their associated coral species.
Biological communities can be organized into trophic levels that rep-
resent the different ways in which species obtain energy from the envi-
ronment (see Figure 2.8). Primary producers make up the first trophic level.
These organisms obtain their energy directly from the sun via photosyn-
thesis. In terrestrial environments, higher plants, such as flowering plants,
gymnosperms, and ferns, are responsible for most photosynthesis, while
in aquatic environments, seaweeds, single-celled algae, and cyanobacteria
(also called blue-green algae) are the most important. All of these species
use solar energy to build the organic molecules they need to live and
grow. Because less energy is transferred to each successive trophic level,
the greatest biomass (living weight) in a terrestrial ecosystem is usually
that of the plants.
The second trophic level contains the herbivores, which eat primary pro-
ducers and are thus known as primary consumers. The intensity of grazing
by herbivores often determines the relative abundance of plant species and
even the amount of plant material present.
Carnivores are in the third and higher trophic levels. Carnivores are
animals that obtain energy by eating other animals. At the third trophic
level are secondary consumers (e.g., foxes), predators that eat herbivores
(e.g., rabbits). At the fourth trophic level are tertiary consumers (e.g., bass),
predators that eat other predators (e.g., frogs).
Some secondary and higher consumers combine direct predation with
scavenging behavior. Others, known as omnivores, include both animal
What Is Biodiversity? 37

and plant foods in their diets. In general, predators occur at lower densi-
ties than their prey, and populations at higher trophic levels contain fewer
individuals than those at lower trophic levels. A single savanna can support
many more zebras than lions.
Parasites and disease-causing organisms, pathogens, form an impor-
tant subclass of predators. Parasites of animals, including mosquitoes, ticks,
intestinal worms, and protozoans, as well as microscopic disease-causing
organisms such as some bacteria and viruses, do not kill their hosts im-
mediately, if ever. Plants can also be attacked by bacteria, viruses, and a
variety of parasites that include fungi, other plants (such as mistletoe),
nematode worms, and insects. The effects of parasites range from imper-
ceptibly weakening their hosts to totally debilitating or killing them over
time. The spread of parasites and disease from captive or domesticated
species, such as dogs, to wild species, such as lions, is a major threat to
many rare species (see Chapter 4).
Decomposers and detritivores feed on dead plant and animal tissues
and wastes (detritus), breaking down complex tissues and organic mol-
ecules into the simple chemicals that are the building blocks of primary
production. Decomposers release minerals such as nitrates and phosphates
back into the soil and water, where they can be taken up again by plants
and algae. Decomposers are usually much less conspicuous than herbivores
and carnivores, but their role in the ecological community is vital. The
most important decomposers are fungi and bacteria, but a wide range of
other species play a role in breaking down organic materials. For example,
vultures and other scavengers tear apart and feed on dead animals, dung
beetles feed on and bury animal dung, and worms break down fallen
leaves and other organic matter. Crabs, worms, molluscs, fish, and numer-
ous other organisms eat detritus in aquatic environments. If decompos-
ers were to die off, organic material would accumulate and plant growth
would decline greatly (Gessner et al. 2010).

Food chains and food webs


Although species can be organized into the general trophic levels we have
just described, their actual requirements or feeding habits within those
trophic levels may be quite restricted. For example, a certain aphid spe-
cies may feed on only one type of plant, and a certain lady beetle species
may feed on only one type of aphid. These specific feeding relationships
are termed food chains. The more common situation in many biological
communities, however, is for one species to feed on several other species at
the lower trophic level, to compete for food with several species at its own
trophic level, and in turn, to be preyed on by several species at the higher
trophic level. Consequently, a more accurate description of the organiza-
tion of biological communities is a food web, in which species are linked
together through complex feeding relationships (Yodzis 2001) (Figure
2.10). Species at the same trophic level that use approximately the same
environmental resources are considered to be a guild of competing species.
38 Chapter 2

Harvest rice

Insects
Rice Eat
ducks
Feed Harvest
Weeds Fixing fish
atmospheric
nitrogen
Duck
manure
Loaches
Azolla
(aquatic plant)
Water fleas,
plankton, worms Manure

FIGURE 2.10 A simple food web in a traditional agricultural ecosystem. Pho-


tosynthetic plants are eaten by people, ducks, and insects. Insects and aquatic
invertebrates are eaten by ducks and fish, which are then eaten by people.

Humans can substantially alter the relationships in food webs (Alva-


Basurto and Arias-Gonzalez 2014). For example, in urban settings, bird
populations may increase due to reduced numbers of predators, reducing
insect abundance in the process (Faeth et al. 2005).

Keystone species and resources


Within biological communities, certain species or guilds of species with
similar ecological features may determine the ability of many other spe-
cies to persist in the community (Figure 2.11). These keystone species
affect the organization of the community to a far greater degree
than we would predict if we considered only their numbers or
biomass (Estes et al. 2011). Protecting keystone species is a pri-
Introduction to Conservation Biology 1E Primack/Sher
Keystone
Sinauer species strongly
Associates ority for conservation efforts because loss of a keystone species
Morales Studio/SA
affect the abundance or guild will lead to losses of numerous other species as well.
Primack_Sher1E_02.10 Date 11-09-15
and distribution of other Top predators are often considered keystone species because
species in an ecosystem. they can markedly influence herbivore populations (Ripple et
Protecting and restoring al. 2014). The elimination of even a small number of individual
keystone species is a predators, even though they constitute only a minute fraction
conservation priority. of the community biomass, may result in dramatic changes in
the vegetation and a great loss in biodiversity, sometimes called
What Is Biodiversity? 39

High
Keystone species Dominant species
(wolves, bats, fig trees, (forest trees, deer,
disease-causing organisms) giant kelps, prairie
grass)
Impact of species

Rare species Common species


(wildflowers, (understory trees,
butterflies, shrubs, weedy
mosses) grasses)
Low
Low Proportional biomass of species High

FIGURE 2.11 Keystone species determine the ability of large numbers of


other species to persist within a biological community. Although keystone spe-
cies make up only a small percentage of the total biomass, a community’s com-
position would change radically if one of them were to disappear. Rare species
have minimal biomass and seldom have significant effects on the community.
Dominant species constitute a large percentage of the biomass and affect many
other species in proportion to this large biomass. Some species, however, have
a relatively low impact on the community organization despite being both com-
mon and heavy in biomass. (After Power et al. 1996.)

a trophic cascade (Jorge et al. 2013; Ripple and Beschta 2012). For example,
in some places where gray wolves (Canis lupis) have been hunted to extinc-
tion by humans, deer (Odocoileus virginiana) populations have exploded.
The deer severely overgraze the habitat, eliminating many herb and shrub
species. The loss of these plants, in turn, is detrimental to the deer and to
other herbivores, including insects. The reduced plant cover may lead to
uction to Conservation Biology 1E Primack/Sher
r Associates soil erosion, also contributing to the loss of species that inhabit the soil.
es Studio/SA When wolves are restored to ecosystems, trophic relationships can some-
ck_Sher1E_02.11 times be Date 11-09-15 1/4/16
reestablished 1/6/16
(Beyer et al. 2007).
Species that extensively modify the physical environment through
their activities, often termed ecosystem engineers, are also considered
keystone species (Jones et al. 1996; Romero et al. 2015) (Figure 2.12). Los-
ing keystone species can create a series of linked extinction events, known
as an extinction cascade, resulting in a degraded ecosystem with much
40 Chapter 2

FIGURE 2.12 The North American beaver (Castor canadensis) is considered


an ecosystem engineer because the dams it constructs cause overbank flooding,
thereby creating new wetland habitat for themselves and other species. (Dam
photo, © Richard Hamilton Smith/Corbis; beaver photo, © iStock.com/Musat.)

lower biodiversity at all trophic levels. This may already be happening in


tropical forests where overharvesting has drastically reduced the popula-
tions of birds and mammals that act as predators, seed dispersers, and
herbivores (Naniwadekar et al. 2015). While such a forest appears to be
green and healthy at first glance, it is really an “empty forest” in which
ecological processes have been irreversibly altered such that many plant
and animal species will be eliminated over succeeding decades or centuries
(Hollings et al. 2014; Redford 1992). In the marine environment, the loss
of key structural species such as sea grasses and seaweeds can lead to the
loss of specialized species that inhabit such communities, such as delicate
sea dragons and sea horses (Hughes et al. 2009). If the few keystone spe-
cies in a community being affected by human activities can be identified,
they can sometimes be carefully protected or even actively managed to
increase their numbers.
Particular habitats may contain keystone resources, often physical or
structural, that occupy only a small area yet are crucial to many species
in the ecosystem (Kelm et al. 2008). For example, deep pools in streams,
What Is Biodiversity? 41

springs, and ponds may be the only refuge for fish and other aquatic
species during the dry season, when water levels drop. For terrestrial
animals, these water sources may provide the only available drinking
water for a considerable distance. Hollow tree trunks and tree holes are
keystone resources as breeding sites for many bird and mammal species
and may limit their population sizes (Cockle and Martin 2015). Protect-
ing old hollow trees as a keystone resource is a priority during certain
logging activities.

Ecosystem dynamics
An ecosystem in which the processes are functioning normally, whether
or not there are human influences, is referred to as a healthy ecosystem.
In many cases, ecosystems that have lost some of their species will remain
healthy because there is often some redundancy in the roles performed
by ecologically similar species. Ecosystems that are able to remain in the
same state are referred to as stable ecosystems. These systems remain
stable either because of lack of disturbance or because they have special
features that allow them to remain stable in the face of disturbance. Such
stability despite disturbance can result from one or both of two features:
resistance and resilience. Resistance is the ability to maintain the same
state even with ongoing disturbance; a river ecosystem that retained its
major ecosystem processes after an oil spill would be considered resis-
tant. Resilience is the ability to return to an original state quickly after
disturbance has occurred; that would be true if, following contamina-
tion by an oil spill and the deaths of many animals and plants, a river
ecosystem soon returned to its original condition (Bhagwat et al. 2012;
Zolli and Healy 2012). As another example, when nonnative fish are intro-
duced into previously fish-free ponds, the number of native animal spe-
cies declines, indicating low resistance. When the fish die out, however,
the number of native species soon recovers, indicating high resilience
(Knapp et al. 2005).

Biodiversity Worldwide
Developing a strategy for conserving biodiversity requires a firm grasp
of how many species exist on Earth and how those species are distributed
across the planet. The answers to both questions can be complex.

How many species exist worldwide?


At present, about 1.5 million species have been described (Costello 2015)
(Figure 2.13). At least two to three times this number of species (primar-
ily insects and other arthropods in the tropics) remain undescribed. Our
knowledge of species numbers is imprecise because inconspicuous species
have not received their proper share of taxonomic attention. For example,
spiders, nematodes (microscopic worms), and fungi living in the soil and
insects living in the tropical forest canopy are small and difficult to study.
42 Chapter 2

FIGURE 2.13 (A) Approximately 1.73 mil-


lion species have been identified and de-
scribed by scientists; the majority of these
species are insects and plants. (B) For those
groups estimated to contain over 100,000
(A) species, the numbers of described species are
indicated by the blue portions of the bars; the
green portions are estimates of the number
Insects 1,000,000 of undescribed species. The vertebrates are
included for comparison. The number of unde-
scribed species is particularly speculative for
the microorganisms (viruses, bacteria, pro-
tists). Most estimates of the possible number
All other of identifiable species range from 5 million to
animals 10 million. (Data from Blackwell 2011, eol.org/
281,000
info/458, Guiry 2012, Pawlowski et al. 2012,
Whitman et al., and Wilson 2010.)

Viruses 5000
Plants 310,442 Bacteria and similar
forms 12,240
Protists (single-celled
organisms with nuclei) Fungi 97,330
43,000 Algae 33,250

Accuracy
(B) of estimate

Viruses Undescribed species Very poor


Described species
Bacteria Very poor

Nematodes Poor

Crustaceans Moderate

Protists Very poor

Algae Very poor

Vertebrates Good

Molluscs Moderate
5 million
Fungi Moderate

Spiders Moderate

Plants Good
5–10 million
Insects Moderate

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8


Number of species (millions)
What Is Biodiversity? 43

These poorly known groups could number in the hundreds of


thousands, or even millions, of species. Our best estimate is Many scientists are
that there are between 5 million and 10 million species on Earth working to determine
(Strain 2011). This number could be many times higher, even up the number of species on
to 100 million species, if it turns out that each species or genus of Earth. The best estimate
animal and plant has many unique species of bacteria, protists, is that there are about
and fungi living on or inside it. 5 million to 10 million
Amazingly, about 20,000 new species are described each species, about half of
year. While certain groups of organisms, such as birds, mam- them insects.
mals, and temperate flowering plants, are relatively well known,
a small but steady number of new species in these groups are
being discovered each year (Joppa et al. 2011). Even among a
group as well studied as primates, dozens of new monkey species have
been found in Brazil, and dozens of new species of lemurs have been
discovered in Madagascar—all since 1990. Every decade, 500–600 new
species of amphibians are described.
Most of the new animals discovered are insects and other invertebrates;
for example, in June 2015, a group of taxonomists and other volunteers dis-
covered seven new species of spiders, including a new genus of tarantula,
in Judbarra/Gregory National Park in Australia’s Northern Territory dur-
ing a one-day event called the Bush Blitz. In a bioblitz, scientists perform
an intensive biological survey of a designated area in a short time with the
goal of documenting all living species there (Figure 2.14). The Bush Blitz
is a bioblitz supported by a partnership between the Australian govern-
ment, museums, and other nonprofit organizations, and since 2010 it has
discovered more than 900 new species, including insects, plants, fish, birds,
and many other taxonomic groups.

FIGURE 2.14 Three genera-


tions of the Dearborn family look
for insect specimens to col-
lect. Acadia National Park, 2009
BioBlitz. (Courtesy of the National
Park Service.)
44 Chapter 2

In addition to new species, entire biological communities continue


to be discovered, often in extremely remote and inaccessible localities.
These communities often consist of inconspicuous species, such as bac-
teria, protists, and small invertebrates, that have escaped the attention
of earlier taxonomists. Specialized exploration techniques have aided in
these discoveries, particularly in the deep sea and in the forest canopy. For
example, each new deep-sea hydrothermal vent explored reveals dozens,
if not hundreds, of species previously unknown to science (Scheckenback
et al. 2010). Drilling projects have shown that diverse bacterial communi-
ties exist 2.8 km deep in the Earth’s crust at densities of up to 100 million
bacteria per gram of solid rock (Fisk et al. 1998). In the depths of the north-
western Pacific near Japan, 50 new species of nematodes were found on
a single collecting trip (Fadeeva et al. 2015). The unique features of these
discoveries not only expand our understanding of the biodiversity of our
planet, but also raise questions regarding evolutionary and physiological
processes (Nakamura and Takai 2015). These organisms are also being
actively investigated as sources of novel chemicals, as medicines, for their
potential usefulness in degrading toxic chemicals, and for insight into
whether life could exist on other planets.
The bacteria in particular are very poorly known (Azam and Worden
2004) and are thus underrepresented in estimates of the total species on
Earth (Dvořák et al. 2015), even as new advances in technology provide
an ever more accurate count of species in this group (Zheng et al. 2015).
Yet they have an important role to play in ecosystem functioning. Only
about 11,000 species of bacteria are currently recognized by microbiolo-
gists (Kyrpides et al. 2014) because they are difficult to grow and identify.
However, analyses of bacterial DNA indicate that there may be from 6400
to 38,000 species in a single gram of soil and 160 species in a milliliter of
seawater (Nee 2003). Such high diversity in small samples suggests that
there could be tens of thousands, or even millions, of undescribed bacterial
species. The Human Microbiome Project has demonstrated that the human
body is occupied by 10 times more bacterial cells than human cells, and
that many still undescribed species occupy specific places both inside our
bodies and on our skin. Furthermore, the diversity and abundance of spe-
cies in different parts of the body is related to the health of an individual
(Human Microbiome Consortium 2012). The human microbiome
is a very active area of current research.
Considering that about 16,000 new animal species are de-
DNA analyses suggest
scribed each year and that at least 3 million more are waiting to
that many thousands of be identified, the task of describing the world’s species will not
species of bacteria have be completed for over 180 years if continued at the present rate.
yet to be described. The This fact underlines the critical need for taxonomists trained
marine environment also to use the latest molecular technology and for web-based in-
contains large numbers formation sharing. International databases such as those of the
of species unknown to Global Biodiversity Information Facility (www.gbif.org) and the
science. Encyclopedia of Life project (www.eol.org) will make species
names and descriptions more widely and readily available.
What Is Biodiversity? 45

Where is the world’s biodiversity found?


The most species-rich environments appear to be tropical forests,
coral reefs, the deep sea, large tropical lakes and river systems, Species diversity is
and regions with Mediterranean climates (Forzza et al. 2012). greatest in the tropics,
particularly in tropical
Tropical forests Even though the world’s tropical forests forests and coral reefs.
occupy only 7% of the land area, they contain more than half the
world’s species, most of which are insects (Corlett and Primack
2010; Gaston and Spicer 2004). Tropical forests also have many
species of birds, mammals, amphibians, and plants. Among flowering
plants, gymnosperms, and ferns, about 40% of the world’s 275,000 species
occur in tropical forest areas. Each of the rain forest areas in the Americas,
Africa, Madagascar, Southeast Asia, New Guinea, Australia, and various
tropical islands has a different biogeographic history, resulting in unique
assemblages of species (see Figure 4.5). For example, lemurs are found
only in Madagascar, and hummingbirds are found only in the Americas.

Oceanic diversity In the oceans, diversity is spread over a much


broader range of phyla and classes than in terrestrial ecosystems. These
marine systems contain representatives of 28 of the 35 animal phyla that
exist today; one-third of these phyla exist only in the marine environment
(Grassle 2001). In contrast, only one phylum is found exclusively in the
terrestrial environment. The broad diversity in the ocean may be due to
its great age, enormous water volume, degree of isolation of some seas by
intervening landmasses, and other factors (Figure 2.15).
Within oceans, coral reef ecosystems are particularly diverse; although
they occupy less than 0.1% of ocean surface area, they are home to one-third
of marine fish species (Bowen et al. 2013; Fisher et al. 2015; see the chapter
opening photo). The physical structure is created by the corals (invertebrate

FIGURE 2.15 A community


of phytoplankton, including
various species of both diatoms
and dinoflagellates. Micro-
scopic organisms such as these
are generally less understood
than those more easily seen,
and new species are frequently
discovered. (© FLPA/Alamy
Stock Photo.)
46 Chapter 2

animals), which provide habitat for many other organisms. The photosyn-
thetic algae that live mutualistically inside the corals provide them with
abundant carbohydrates. One explanation for the richness of coral reefs
is their high primary productivity: 2500 grams of biomass (living matter)
per square meter per year, in comparison with 125 g/m 2/y in the open
ocean. Extensive niche specialization among coral species and adaptations
to varying levels of disturbance may also account for the high species rich-
ness found in coral reefs. The world’s largest coral reef is Australia’s Great
Barrier Reef, with an area of 349,000 km2, which contains over 400 species
of corals, 1500 species of fish, 4000 species of molluscs, and 6 species of
turtles. Coral reefs also support some 252 species of birds.

Mediterranean-type communities Great diversity is found among


plant species in Southwest Australia, the Cape region of South Africa,
California, central Chile, and the Mediterranean basin, all of which are
characterized by a Mediterranean climate of moist winters and hot, dry
summers (see Figure 4.5). Of these regions, the Mediterranean basin is the
largest in area (2.1 million km 2) and has the most plant species (22,500)
(Conservation International and Caley 2008). The Cape Floristic Region of
South Africa has an extraordinary concentration of unique plant species
(9000) in a relatively small area (78,555 km2) (Figure 2.16). The shrub and
herb communities in these areas are rich in species, apparently because of
their combination of considerable geologic age, complex site characteristics
(such as topography and soils), and severe environmental conditions. The
high frequency of fire in these areas may also favor rapid speciation and
prevent the domination of just a few species.

FIGURE 2.16 The Cape


region of South Africa has
evolved a unique ecosys-
tem, collectively called the
fynbos (meaning “fine-leaved
plants” in Dutch), charac-
terized by large numbers of
plant species that are found
nowhere else. (Photograph
by Anna Sher.)
What Is Biodiversity? 47

The distribution of species


At various spatial scales, there are concentrations of species in particular
places, and there is a rough correspondence in the distribution of species
richness between different groups of organisms (Domisch et al. 2015; Si-
maika et al. 2013). For example, in North America, large-scale patterns of
species richness are highly correlated for amphibians, birds, butterflies,
mammals, reptiles, land snails, trees, all vascular plants, and tiger beetles;
that is, a region with numerous species of one group will tend to have nu-
merous species of the other groups (Ricketts et al. 1999). On a local scale,
however, this relationship may break down; for example, amphibians may
be most diverse in wet, shady habitats, whereas reptiles may be most di-
verse in drier, open habitats. At a global scale, each group of living organ-
isms may reach its greatest species richness in a different part of the world
because of historical circumstances or the suitability of the site to its needs.
Places with large concentrations of species often have a high percentage
of endemic species, that is, species that occur there and nowhere else. The
countries with the largest numbers of endemic mammals, which represent
important targets for conservation efforts, are Indonesia (259), Australia
(241), Brazil (183), Madagascar (187), and Mexico (158) (www.iucnredlist.
org). Geographically isolated countries and islands also tend to have high
percentages of endemic species. For example, most of the species in Mada-
gascar are endemic because it is a large, ancient island on which many
unique species have evolved in isolation. Species-rich countries on the
mainland, such as Tanzania, have comparatively fewer endemic species
because they share many species with neighboring countries.
Almost all groups of organisms show an increase in species diversity
toward the tropics (Groombridge and Jenkins 2010) (see Figures 4.5 and 6.20).
For example, Thailand has 241 species of mammals, while France has only
104, despite the fact that both countries have roughly the same land area
(Table 2.1). The contrast is particularly striking for trees and other flowering
plants: 10 hectares of forest in Amazonian Peru or Brazil might have 300 or
more tree species, whereas an equivalent forest area in temperate Europe or
the United States would probably contain 30 species or less. Within a given
continent, the number of species increases toward the equator.
For many groups of marine species, the greatest diversity of coastal
species is found in the tropics, with a particular richness of species in the
western Pacific. For open-ocean species, the greatest diversity is found at
midlatitudes. Temperature is the most important variable explaining these
patterns (Tittensor et al. 2010).
Local variation in climate, sunlight and rainfall, topography, and geo-
logic age also affects patterns of species richness (Zellweger et al. 2015). In
terrestrial communities, species richness tends to increase with decreas-
ing elevation, increasing solar radiation, and increasing precipitation; that
is, hot, rainy lowland areas have the most species. Species richness can be
greater where complex topography and great geologic age provide more en-
vironmental variation, which allows genetic isolation, local adaptation, and
speciation to occur. Geologically complex areas can produce a variety of soil
48 Chapter 2

Table 2.1 Number of Native Mammal Species in Selected


Tropical and Temperate Countries Paired for
Comparable Size
Number of Number of
Tropical Area mammal Temperate Area mammal
country (1000 km2) species country (1000 km2) species
Brazil 8456 648 Canada 9220 202
DRCa 2268 430 Argentina 2737 374
Mexico 1909 523 Algeria 2382 105
Indonesia 1812 670 Iran 1636 186
Colombia 1039 442 South 1221 297
Africa
Venezuela 882 363 Chile 748 143
Thailand 511 311 France 550 123
Philippines 298 207 United 242 74
Kingdom
Rwanda 25 184 Belgium 30 70
Source: Data from 2008 IUCN Red List, accessed in 2015.
a
DRC = Democratic Republic of the Congo.

conditions with very sharp boundaries between them, leading to multiple


communities and plant species adapted to one specific soil type or another.
With better methods of exploration and investigation, we are now able
to appreciate the great diversity of the living world. This is truly a golden
age of biological exploration. Natural history societies and clubs that com-
bine amateur and professional naturalists contribute to this effort. Yet with
this knowledge of biodiversity comes both the awareness of the damaging
impact of human activity, which is diminishing it right before our eyes, and
the responsibility to protect and restore that biodiversity that still remains.

Summary
„„Taxonomists use morphological and the continuing improvement of crop
genetic information to describe and plants and domesticated animals.
identify the world’s species. Places vary „„Within an ecosystem, species play dif-
in their species richness, the number of ferent roles and have varying require-
species found in a particular location. ments for survival. Certain keystone
„„There is genetic variation among indi- species are important in determining the
viduals within a species. Genetic varia- ability of other species to persist in an
tion allows species to adapt to a chang- ecosystem.
ing environment, and it is valuable for
What Is Biodiversity? 49

„„It is estimated that there are 5 million to „„The greatest biological diversity is found
10 million species, most of which are in- in tropical regions, with particular con-
sects. The majority of the world’s species centrations of species in rain forests and
have still not been described and named. coral reefs. The ocean may also have
Further work is needed to describe mi- great species diversity but needs further
croorganisms such as bacteria. exploration.

For Discussion
1. How many species of birds, trees, and in- 3. Conservation efforts usually target ge-
sects can you identify in your neighbor- netic variation, species diversity, bio-
hood? How could you learn to identify logical communities, and ecosystems for
more? Is it important to be able to iden- protection. What are some other com-
tify species in the wild? ponents of natural systems that need to
2. What are the factors promoting species be protected? What do you think is the
richness? Why is biological diversity most important component of biodiver-
diminished in particular environments? sity, and why do you believe it is most
Why aren’t species able to overcome important?
these limitations and undergo the process
of speciation?

Suggested Readings
Albert, A., K. McKonkey, T. Savini, and M. C. Huynen. 2014. The value of dis-
turbance-tolerant cercopithecine monkeys as seed dispersers in degraded
habitats. Biological Conservation 170: 300–310. Monkeys are important in
dispersing seeds and helping the forest to regenerate.
Chan, Y. F., K.-P. Chiang, J. Chang, Ø. Moestrup, and C.-C. Chung. 2015.
Strains of the morphospecies Ploeotia costata (Euglenozoa) isolated from
the Western North Pacific (Taiwan) reveal substantial genetic differences.
Journal of Eukaryotic Microbiology 62: 318–326.
Corlett, R. and R. B. Primack. 2010. Tropical Rainforests: An Ecological and Bio-
geographical Comparison, 2nd Ed. Wiley-Blackwell Publishing, Malden, MA.
Rain forests on different continents have distinctive assemblages of animal
and plant species.
González-Maya, J. F., L.R. Víquez-R, A. Arias-Alzate, J.L. Belant, and G. Cebal-
los. 2016. Spatial patterns of species richness and functional diversity in
Costa Rican terrestrial mammals: implications for conservation. Diversity
and Distributions 22: 43–56. Understanding the relationship between differ-
ent measures of diversity can help us understand ecosystem function.
Groombridge, B. and M. D. Jenkins. 2010. World Atlas of Biodiversity: Earth’s
Living Resources in the 21st Century. University of California Press, Berkeley.
Great resource, with numerous figures; available on-line.
Hollings, T., M. Jones, N. Mooney, and H. McCallum. 2014. Trophic cascades
following the disease-induced decline of an apex predator, the Tasmanian
devil. Conservation Biology 28: 63–75. Many species are affected when a
keystone species is eliminated.
50 Chapter 2

Joppa, L. N., D. L. Roberts, and S. L. Pimm. 2011. The population ecology and
social behavior of taxonomists. Trends in Ecology and Evolution 26: 551–553.
The number of taxonomists and the number of species described per year
are steadily increasing.
Laikre, L. and 19 others. 2010. Neglect of genetic diversity in implementation
of the Convention on Biological Diversity. Conservation Biology 24: 86–88.
A greater emphasis on genetic diversity needs to be part of conservation
efforts.
Magurran, A. E. 2013. Measuring Biological Diversity: Frontiers in Measurement
and Assessment. Oxford University Press, Oxford, UK. A widely-cited text
that discusses both classic and emerging methods.
Ricklefs, R. E., and F. He. 2016. Region effects influence local tree species di-
versity. Proceedings of the National Academy of Sciences USA, 113: 674–679.
Regional species diversity results largely from geologic and geographic
properties that affect evolution.
Ripple, W. J. and R. L. Beschta. 2012. Trophic cascades in Yellowstone: The first
15 years after wolf reintroduction. Biological Conservation 145: 205–213. Re-
storing a keystone species has resulted in large changes to this ecosystem.
Strain, D. 2011. 8.7 million: A new estimate for all the complex species on
Earth. Science 333: 1083. A variety of methods have been developed for
estimating the total numbers of species on Earth.
Tittensor, D. P. and 6 others. 2010. Global patterns and predictors of marine
biodiversity across taxa. Nature 466: 1098–1101. Temperature is the most
important factor affecting marine diversity.
The Value of
3 Biodiversity
Ecological and Existence Value 80
Environmental Economics 55 Environmental Ethics 83
Economic Use Values 61
The Long-Term View:
Option Value 78

Ecotourists viewing Cecil the Lion at


Hwange National Park in Zimbabwe.
I
n July 2015, the economics of biodiversity
received international attention when a lion
from Hwange National Park in Zimbabwe
was illegally shot by an American big-game hunter
who had paid poachers $50,000 to do so (CNN 2015).
The 13-year-old male, affectionately named “Cecil the
Lion,” had been an important animal icon and tour-
ism draw for the park; there was an immediate outcry
from conservationists and animal lovers around the
world when his death was announced. What Cecil
and other wild organisms are “worth,” both alive and
dead, is a critical question in evaluating attitudes and
implementing actions that affect biodiversity.
Ultimately, the trend of biodiversity decline will
be reversed only if people believe that we are truly
losing something of value. But what exactly are we
losing? Why should anyone care if a species becomes
extinct or an ecosystem is destroyed? What factors
induce humans to act in an unsustainable and ulti-
mately destructive manner? What are we willing to
spend to protect species, and how will this spending
be financed? Governments and communities through-
out the world are coming to realize that biodiversity is
extremely valuable—indeed, it is essential to human
existence.
54 Chapter 3

Most environmental degradation and species losses occur as acci-


dental by-products of human activities. Species are hunted to extinction.
Sewage is released into rivers. Low-quality land is cleared for short-
term cultivation. Economics, the study of the transfer of the production,
distribution, and consumption of goods and services, can both help us
understand the reasons why people treat the environment in what ap-
pears to be a shortsighted manner and provide tools to help protect en-
vironmental resources.
One of the most universally accepted tenets of modern economic
thought centers on the “voluntary transaction”: the idea that a monetary
transaction takes place only when it is beneficial to both of the parties
involved. For example, a baker who sells his loaves of bread for $40 will
find few customers. Likewise, a customer who is willing to pay only 4 cents
for a loaf will soon go hungry. A transaction between seller and buyer will
occur only when a mutually agreeable price is set that benefits both par-
ties: perhaps $4 for that loaf of bread. Adam Smith, the eighteenth-century
philosopher whose ideas are the foundation of much modern economic
thought, wrote, “It is not upon the benevolence of the butcher, the baker,
or the brewer that we eat our daily bread, but upon his own self-interest”
(Smith 1909). The sum of each individual’s actions in his or her self-interest
results in a more prosperous society.
There are exceptions to Smith’s principle, however, that apply directly
to environmental issues. For example, Smith assumed that all the costs and
benefits of free exchange are accepted and borne by the participants in the
transaction. In some cases, however, associated costs (or sometimes benefits)
befall individuals not directly involved in the exchange. These hidden costs
and benefits are known as negative and positive externalities, respectively
(Figure 3.1) (Abson and Termansen 2011). Companies, individuals, or other
stakeholders involved in production that results in ecological damage gener-
ally do not bear the full cost of their activities, but gain substantial private
economic benefits. For example, the company that owns an electric power
plant that burns coal and emits toxic fumes benefits from the sale of low-cost
electricity, as does the consumer who buys this electricity. Yet the nega-
tive externalities of this transaction—decreased air quality and visibility,
increased respiratory disease for people and animals, damage to plant life,
and a polluted environment—are distributed throughout society.
Even more important and more frequently overlooked is the negative
externality of environmental damage done to open-access resources, such
as water, air, and soil, as a consequence of human economic activity. Open-
access resources are those that are collectively owned by society at large
(also called common-property resources) or owned by no one. They are
available for everyone to use and are essentially free. When there are no
regulations on their use, then people, industries, and governments use
and damage these resources without paying more than a minimal cost,
or sometimes nothing at all. This situation, which has been referred to
as the tragedy of the commons (a phrase popularized by Garett Hardin’s
1968 essay of that name on human population control), means that the
The Value of Biodiversity 55

FIGURE 3.1 Smog in the Andes, looking eastward over Santiago, Chile. Air
pollution is responsible for 3.3 million premature human deaths (and unknown
numbers of wildlife) worldwide per year, a devastating externality of the combined
emissions from industry, transportation, and agriculture. (© Matt Mawson/Corbis.)

value of the open-access resource is gradually lost to all of society. The


unregulated dumping of industrial sewage into a river as a by-product of
manufacturing is a common example. The externalities of this activity are
degraded drinking water and an increase in disease, loss of opportunity
to bathe and swim in the water, fewer fish that are safe to eat, and the loss
of many species unable to survive in the polluted river. The factory owner
gains free disposal of sewage, but the society pays the price in terms of
lost products and services.
When externalities are not identified and managed, or there is inade-
quate regulation of common-property resources, certain economic activities
make the society as a whole less prosperous, not more prosperous. When
an economic system fails, and thus the balance of supply and demand for
products or services is lost, economists call this market failure (Figure
3.2). Avoiding market failure is arguably the primary goal of government
policies that regulate economic activities that can affect the environment.

Ecological and Environmental Economics


A major problem for conservation biology is that natural resources have
often been undervalued by modern society. Thus, the costs of environmental
damage have been ignored, the depletion of natural resource stocks disre-
56 Chapter 3

Ecosystem services preserved Ecosystem services damaged Intensive


Wildlife refuge grazing

Clear-cutting and erosion Erosion


Selective logging gully
Mining
Water tower Intensive
hway agriculture
Light Hig
grazing Moun Town
tain Town
ro Siltation
ad

r
rive
Water treatment Stream carrying
Low-intensity
ve r

ted
plant Road untreated wastewater Feedlot
farming
r ri

llu
Factory
ate

Po
Wildlife refuge
n-w

Beach resort
Clea

Danger! Siltation
Poisons! plume Polluted
Coastal fishing No fishing! runoff

Figure 3.2 Agricultural ecosystems, forestry activities, and industries are


usually valued by the products that they produce. In many cases, these activities
have negative externalities in that they erode soil, degrade water quality, and
harm aquatic life (right side of figure). Farming, forestry, and other human activi-
ties could also be valued on the basis of their public benefits, such as soil reten-
tion and maintaining water quality and fish populations, and their owners might
receive subsidies for these benefits (left side of figure).

garded, and the future value of resources discounted (MEA 2005). Because
the underlying causes of environmental damage are so often economic in
nature, the solution must incorporate economic principles (Kubiszewski
et al. 2013). In an effort to account for all costs of economic transactions,
including environmental costs, two closely related research areas have
evolved—environmental economics and ecological economics—that inte-
grate economics, environmental science, ecology, and public policy and that
include valuations of biodiversity in economic analyses (Common and Stagl
2005). Environmental economics is a subdiscipline of economics that places
Introduction to Conservation Biology 1E Primack/Sher
Sinauer Associates a value on components of the environment. Its modern form can
Morales Studio/SA be traced to a popular 1972 book, The Limits to Growth, by envi-
Primack_Sher1E_03.02 Date 11-24-15
ronmental scientist Donella Meadows and coauthors; they used
Arguments for the system dynamics modeling to predict potential future balances
protection of biodiversity of the human population, pollution, and agricultural production
are often strengthened (Hanley et al. 2015). Ecological economics, which is more closely
by evidence provided by allied to conservation biology, seeks to integrate the thinking of
ecological economics. ecologists and economists into a transdiscipline aimed at devel-
oping a sustainable world (Sachs 2008). One of the core agenda
The Value of Biodiversity 57

items of ecological economics is to develop methods to value biodiversity


by integrating economic valuation with ecology, environmental science, so-
ciology, and ethics and to use those new valuation methods to design better
public policies related to conservation and environmental issues (Reyers
et al. 2013). The fundamental challenge facing conservation biologists is to
ensure that all the costs of economic activity, as well as all the benefits, are
understood and taken into account when decisions that will affect biodi-
versity are made (Hoeinghaus et al. 2009; Junk et al. 2014).

Cost–benefit analysis
Economic methods are now being used to review development projects and
evaluate their potential environmental effects before the projects proceed.
Environmental impact assessments, in particular, consider the present
and future effects of projects on the environment. “The environment” is
often broadly defined to include not only harvestable natural resources
but also air and water quality, the quality of life for local people, and bio-
diversity. In its most comprehensive form, cost–benefit analysis compares
the values gained against the costs of a project or resource use (Maron et
al. 2013; Newbold and Siikamäki 2009). In practice, though, cost–benefit
analyses are notoriously difficult to calculate accurately because benefits
and costs change over time and are difficult to measure. Today, there is
an increasing tendency by governments, conservation groups, and econo-
mists to apply the precautionary principle. That is, it may be better not to
approve a project that has risk associated with it and to err on the side of
doing no harm to the environment, rather than doing harm unintention-
ally or unexpectedly, as by building wind turbines where they could harm
endangered birds (Braunisch et al. 2015). The precautionary principle is a
key feature of many national and international policies and agreements
regarding environmental management, even though its interpretation can
be vague and variable (Foster et al. 2000).
It would be highly beneficial to apply cost–benefit analysis to many of
the basic industries and practices of modern society. Many environment-
damaging economic activities appear to be profitable even when they are ac-
tually losing money because governments subsidize the industries involved
in them with tax breaks, direct payments or price supports, cheap fossil
fuels, free water, and road networks—sometimes referred to as perverse
subsidies (Myers et al. 2007). The elimination of such subsidies that are
harmful to biodiversity by 2020 is one of the explicit targets of the Conven-
tion on Biological Diversity (CBD Decision X/2; Dobson 2005). Subsidies in
agriculture and fisheries can be as high as 20%–30% of the production value
of those industries (MEA 2005). Without these subsidies, many environmen-
tally damaging or expensive activities—such as farming in areas with high
labor, energy, and water costs; overfishing in the ocean; and inefficient and
highly polluting energy use—would be reduced (Merckx and Pereira 2015).
Attempts have been made to include the loss of natural resources in
calculations of gross domestic product (GDP) and other indexes of national
58 Chapter 3

production. The problem with GDP is that it measures economic


Unsustainable activities activity in a country without accounting for all the costs of un-
such as clear-cut logging, sustainable activities (such as overfishing of coastal waters and
strip mining, and over- poorly managed strip mining), which cause the GDP to increase,
fishing may cause a even though they may be destructive to a country’s long-term
country’s apparent economic well-being. In actuality, the economic costs associated
productivity to increase
with environmental damage can be considerable, and they often
offset the gains attained through agricultural and industrial
temporarily, but are
development. In the United Kingdom, hidden environmental
generally destructive
costs in agriculture, including soil erosion and water pollution,
to long-term economic
are estimated to be worth about $2.6 billion per year, or 9% of
well-being.
the value of the country’s agriculture (MEA 2005).
A system that accounts for natural resource depletion, pollu-
tion, and unequal income distribution in measures of national
production is the Index of Sustainable Economic Welfare (ISEW),
the updated version of which is called the Genuine Progress Indicator (GPI;
www.progress.org). This index includes factors such as the loss of farm-
lands, the loss of wetlands, the impact of acid rain, the number of people
living in poverty, and the effects of pollution on human health. According
to the GPI, the world economy reached a peak around 1978 and has been
slowly declining since then, even though the standard GDP index showed a
dramatic gain (Kubiszewski et al. 2013). The GPI suggests what conservation
biologists have long feared: Many modern economies are achieving their
growth only through the unsustainable consumption of natural resources
and environmental degradation. As these resources run out and as humans
suffer the effects of pollution, the true economic situation will continue to
deteriorate.
A third measure of national productivity is the Environmental Per-
formance Index (EPI), which uses 20 environmental indicators to rank
countries according to the health of, and threats to, their ecosystems; the
vulnerability of their human population to adverse environmental condi-
tions; the ability of their society to protect the environment; and their
participation in global environmental protection efforts (epi.yale.edu). In
general, developing countries with a low GDP per person also
have low EPI scores, including such large countries as China
and India, as Figure 3.3 shows. Higher-income countries, such
as the United States, Japan, and Germany, tend to have much
New measures of higher EPI scores. There is a concern among many economists
national productivity and businesspeople that a country that rigorously protects its
take environmental environment, as shown by a high EPI, might not be competi-
sustainability into tive in the world economy, as measured by a competitiveness
account. These measures index that includes worker productivity and a country’s abil-
include both the benefits ity to grow and prosper. But environmental sustainability is
and the costs of human not linked to a country’s economic competitiveness. Countries
activities. such as Finland have an economy that is both sustainable and
competitive, whereas Belgium is competitive but ranks poorly in
The Value of Biodiversity 59

120
Norway
100
(thousands of U.S. dollars)

Switzerland
80 Australia
GDP per person

United States Japan


60
Germany
40
China
India
20
Somalia
0 United
Kingdom

20 30 40 50 60 70 80
EPI score

Figure 3.3 Wealthy countries with higher gross domestic product (GDP)
per person tend to have higher scores on the Environmental Performance Index
(EPI), as measured by various indicators: health and stress level of ecosystems,
human vulnerability to environmental change, ability of the society and insti-
tutions to cope with environmental changes, and cooperation in international
environmental initiatives. The size of the human population is indicated by the
size of the circle. (After epi.yale.edu/epi/data-explorer.)

sustainability. The rapidly growing economies of China and India are inter-
mediate in competitiveness, but rank low in environmental sustainability.

Financing conservation
Another important aspect of environmental economics is the cost of con-
servation. Especially when a species is already rare or endangered, it is
not enough to simply do no harm. People must intervene to protect, man-
age, and otherwise support its health and survival. But such interventions
can be expensive. In an extreme example, whooping crane conservation is
estimated to cost $2–2.5 million per year, with a total projected cost of $48
million (US Fish and Wildlife Service). The effort increased this species
from
o Conservation Biology 1E16Primack/Sher
individuals in 1941 to 603 as of 2015. Although the rescue of the
ciates whooping crane is a success story, there are many other less charismatic
io/SA
r1E_03.03
species that have less financial support than they need. Thus, in addition
Date 11-24-15 11-25-15 1/5/16
to cost–benefit analysis, it is important to do cost-effectiveness analysis
as well, or to ask, “Where do we get the most with our conservation dol-
lar?” (Cannon 1996).
We must also concern ourselves with where and how we get these
funds. Conservation is financed in many ways, including by the some-
times-controversial source of hunting (Crosmary et al. 2015). Some coun-
tries, such as Namibia, depend on sales of expensive trophy licenses, even
for the hunting of endangered species, to support conservation (Rust
2015). In the United States, most state wildlife conservation efforts are
primarily funded by the sale of hunting licenses, tags, and stamps, and
60 Chapter 3

$200 million a year in hunters’ federal excise taxes is distributed to states,


primarily for wildlife management (US Fish and Wildlife Service). This is
considered a utilitarian approach to conservation: it assumes that species
and their habitats will be conserved if people find them valuable, even
if only to kill for sport. In Great Britain, landowners were much more
likely to plant and protect woodland for nonagricultural purposes if
they hunted game there (Oldfield et al. 2003). However, it has also been
argued that using hunting as a foundation for conservation is morally
wrong and will inevitably lead to poaching and decreased biodiversity
(Selier et al. 2014).
Ultimately, only that which is perceived to have value will be saved,
so the basis on which we assign this value is of utmost importance. It will
also inevitably raise questions regarding moral and ethical values.

What are species worth?


There are many classification systems used to evaluate the benefits we
receive from our natural environment (Wallace 2007). As yet there is no
universally accepted framework for assigning value to biodiversity, but a
variety of approaches have been proposed. Among the most useful is the
framework used by McNeely et al. (1990) and Barbier et al. (1994), in which
economic values are first divided into use values and non-use values. Use
values of biodiversity are divided between direct use values (also known
in other frameworks as commodity values and private goods) and indi-
rect use values. Direct use values are assigned to products harvested by
people, such as timber, seafood, and medicinal plants from the wild, while
indirect use values are assigned to benefits provided by biodiversity that
do not involve harvesting or destroying the resource. Indirect use values
provide current benefits to people, such as recreation, education, scientific
research, and scenic amenities, and include the benefits of ecosystem ser-
vices, such as water quality, pollution control, natural pollination and pest
control, ecosystem productivity, soil protection, and regulation of climate.
Option value is determined by the prospect for possible future benefits for
human society, such as new medicines, possible future food sources, and
future genetic resources. Existence value is the non-use value that can be
assigned to biodiversity—for example, economists can attempt to measure
how much people are willing to pay to protect a species from going extinct
or an ecosystem from being destroyed.

Ecosystem services
The many and varied environmental benefits provided by biodiversity and
ecosystems in general to humans are collectively referred to as ecosystem
services (Ehrlich and Ehrlich 1982), which are typically divided into four
categories:
Provisioning services are the material or energy outputs of an ecosystem,
including food, fresh water, and raw materials. The worth of biological
outputs to humans will be discussed in detail below under “Direct use
The Value of Biodiversity 61

values,” whereas non-living products such as clean water give the organ-
isms that supply it indirect economic value.
Regulating services are services provided by the ecosystem acting as
regulators of the quality of the air and soil. Forests provide many of these
by regulating local climate, removing pollutants from the atmosphere, and
holding soil with their roots that would otherwise blow or wash away.
Other examples can be found in the section “Indirect use values.”
Habitat/supporting services refers to the role ecosystems play in support-
ing biodiversity, including genetic diversity that humans depend on for
cultivating crops and livestock. Other examples can be found in “Species
relationships and environmental monitors.” These supportive services
mean that they also provide option use value.
Cultural services include inspiration for art, design, music and other
cultural expression, aesthetics, intellectual stimulation, and spiritual value.
Examples of these can be found under “Amenity value,” “Education and
scientific value,” and “Existence Value.”
These categories were first defined in the Millennium Ecosystem As-
sessment (2005), a project initiated by the United Nations in 2000 to evalu-
ate the impact of ecosystem change on human well-being and determine
which actions were needed to protect these services. Below we will discuss
these services in terms of the value that the organisms (both living and
dead) themselves have for humans and how this value is quantified in
financial terms.

Economic Use Values


Direct use values
Direct use values can often be readily calculated by observing the activi-
ties of representative groups of people, by monitoring collection points for
natural products, and by examining import and export statistics. Direct
use values are further divided into consumptive use value, for goods that
are consumed locally, and productive use value, for products that are sold
in markets.

Consumptive use value


People living close to the land often derive a considerable proportion of the
goods they require for their livelihood from the surrounding environment
(Figure 3.4). These goods, such as fuelwood and wild meat, are consumed
locally and are therefore assigned consumptive use value (Davidar et al.
2008). These goods do not appear in the GDP of countries because they are
neither bought nor sold beyond the village or local region and do not ap-
pear in the national or international marketplace. However, if rural people
are unable to obtain these products (as might occur following environmen-
tal degradation, overexploitation of natural resources, or even creation of
a protected reserve), their standard of living will decline, possibly to the
62 Chapter 3

(A)

(B)

Figure 3.4 Examples of natural


products with consumptive use value:
(A) Women in India return to their village
with loads of wood. Fuelwood is one of the
most important natural products consumed
by local people, particularly in Africa and
southern Asia. The value of these products
can be estimated based on what these
people would have to spend to purchase
rather than harvest them. (B) Along a river
in India, fishermen catch small fish to eat.
(A, photograph © Robert Harding World
Imagery/Corbis; B, photograph courtesy of
Sandesh Kadur.)
The Value of Biodiversity 63

point where they are forced to relocate. Consumptive use value can be as-
signed to a product by considering how much people would have to pay
if they had to buy an equivalent product when their local source was no
longer available. This valuation is sometimes referred to as a replacement
cost approach.
Studies of traditional societies in the developing world show how ex-
tensively these people use their natural environment to supply themselves
with fuelwood (Figure 3.4A), meat, vegetables, fruit, medicine, rope and
string, and building materials (Angelsen et al. 2014). About 80% of the
world’s population still relies principally on traditional medicines derived
from plants and animals as their primary source of treatment (Shanley
and Luz 2003).
One of the crucial requirements of rural people is protein, which they
obtain by hunting and collecting wild animals for meat. In some places,
this meat is called bushmeat. In many areas of Africa, bushmeat constitutes
a significant portion of the protein in the average person’s diet—about
40% in Botswana and about 80% in the Democratic Republic of the Congo
(formerly Zaire; Powell et al. 2013). Bushmeat extraction rates for Africa
are undeniably unsustainable, perhaps by a factor of six. This wild meat
includes not only birds, mammals, and fish, but spiders, snails, caterpil-
lars, and insects. In certain areas of Africa, because of overharvesting of
larger animals, insects may constitute the majority of the dietary protein
and supply critical vitamins.
In areas along coasts, rivers, and lakes, wild fish represent an important
source of protein (Figure 3.4B). Throughout the world, 130 million tons of
fish, crustaceans, and molluscs, mainly wild species, are harvested each
year, 100 million tons from the oceans and 30 million tons from freshwa-
ter (Chivian and Bernstein 2008). Much of this catch is consumed locally.
In coastal areas, fishing is often the most important source of
employment, and seafood is the most widely consumed protein.
Even though fish farming is increasing rapidly, much of the feed
Consumptive use value
used is fish meal derived from wild-caught fish (Gross 2008).
can be calculated by
Although dependency on local natural products is primarily
considering how much
associated with the developing world, there are rural areas of
people would have to
the United States, Canada, Europe, and other developed coun-
tries where hundreds of thousands of people are dependent pay to buy an equivalent
on fuelwood for heating, on wild game and seafood for their product if their local
protein needs, and on intact ecosystems for clean drinking water source were no longer
and sewage treatment. Many of these people would be unable to available.
survive in these locations if they had to pay for these necessities.

Productive use value


Resources that are harvested from the wild and sold in national or inter-
national commercial markets are assigned productive use value (Figure
3.5). In standard economics, these products are valued at the price paid at
the first point of sale minus the costs incurred up to that point, whereas
64 Chapter 3

Figure 3.5 Examples of (A)


productive use value. (A) A
wide variety of marine animals
are collected in the wild or pro-
duced by aquaculture and then
sold as seafood, as shown by
this market in South Korea. (B)
The productive use value of the
trees in this forest in Canada
is simply calculated as their
worth on the market. (A, pho-
tograph by Richard B. Primack;
B, photograph © Lloyd Sutton/
Alamy Stock Photo.)

(B)
The Value of Biodiversity 65

other methods value the resource at the final retail price of the products.
For example, the bark and leaves from wild shrubs and trees of the common
witch hazel (Hamamelis virginiana and related species) are used to make a
variety of astringent herbal products, including aftershave lotions, insect-
bite creams, and hemorrhoid preparations. The final retail price of the medi-
cine, which includes the values of all inputs (labor, energy, other materials,
transportation, and marketing, as well as witch hazel bark and leaves), is
vastly greater than the purchase price of the witch hazel raw materials.
The productive use value of natural resources is significant, even in
industrial nations. It has been estimated that approximately 4.5% of the US
GDP depends in some way on wild species (Prescott-Allen and Prescott-
Allen 1986). This translates to about $780 billion (out of a GDP of $17.4 tril-
lion) for the year 2014. The percentage is far higher for developing countries
that have less industry and a higher percentage of their population living
in rural areas. The international trade in wildlife, fisheries, and timber
products harvested from the wild has been estimated to be $332 billion (En-
gler 2008, cited in Barber-Meyer 2010). However, it is difficult to accurately
calculate the total value of wild-harvested products because of the unknown
contribution of “invisible trades” due to low detection rates, underreporting,
and non-reporting, especially of illegal products (Phelps and Webb 2015).
The range of products obtained from the natural environment and sold
in the marketplace is enormous: these products include fuelwood,
construction timber, fish and shellfish, medicinal plants, wild
fruits and vegetables, wild meat and skins, fibers, rattan (a vine
used to make furniture and other household articles), honey, bees- A wide variety of natural
wax, natural dyes, seaweed, animal fodder, natural perfumes, resources are sold
and plant gums and resins (Baskin 1997; Chivian and Bernstein commercially and have
2008). Additionally, there are large international industries associ- enormous total market
ated with collecting tropical cacti, orchids, and other plants for value. Their value can be
the horticultural industry and birds, mammals, amphibians, and considered the productive
reptiles for zoos and private collections. The value of ornamental value of biodiversity.
fishes in the aquarium trade is estimated at $1 billion per year,
with wild-caught fish representing about 20% of the total.

Forest products Wood is one of the most significant products ob-


tained from natural environments, with an export value of about $231
billion per year (www.fao.org/forestry/statistics/80938). The total value
of timber and other wood products is far greater—perhaps about $400
billion per year—because most wood is used locally and is not exported.
In tropical countries such as Indonesia, Brazil, and Malaysia, timber prod-
ucts earn billions of dollars per year (Corlett and Primack 2010). Non-wood
products from forests, including bushmeat, fruits, gums and resins, rattan,
and medicinal plants, also have a large productive use value. These non-
wood products are sometimes erroneously called “minor forest products”;
in reality, they are often very important economically and may even rival
the value of wood.
66 Chapter 3

The natural pharmacy Effective drugs are needed to keep people


healthy, and they represent an enormous industry, with worldwide sales of
about $300 billion per year (Chivian and Bernstein 2008). The natural world
is an important source of medicines currently in use as well as possible
future medicines. All 20 of the pharmaceutical products most frequently
used in the United States are based on chemicals that were first identified
in natural organisms. More than 25% of the prescriptions filled in the
United States contain active ingredients derived directly from plants, and
many of the most important antibiotics, including penicillin and tetracy-
cline, are derived from fungi or microorganisms (Waterman et al. 2016).
Many modern medicines were first discovered in a wild species
used in traditional medicine, and then produced synthetically by chem-
ists (Chivian and Bernstein 2008; Cox 2001). For example, the use of coca
(Erythroxylum coca) by natives of the Andean highlands eventually led to
the development of synthetic derivatives such as Novocain, procaine, and
lidocaine, commonly used as local anesthetics in dentistry and surgery. The
rose periwinkle (Catharanthus roseus) from Madagascar (Table 3.1) is the

Table 3.1 
Twenty Drugs from the Plant World First Discovered in Traditional
Medical Practice
Drug Medical use Plant source Common name
Ajmaline Treats heart arrhythmia Rauwolfia spp. Rauwolfia
Aspirin Analgesic, anti-inflammatory Spiraea ulmaria Meadowsweet
Atropine Dilates eyes during Atropa belladonna Belladonna
examination
Caffeine Stimulant Camellia sinensis Tea plant
Cocaine Ophthalmic analgesic Erythroxylum coca Coca plant
Codeine Analgesic, antitussive Papaver somniferum Opium poppy
Digitoxin Cardiac stimulant Digitalis purpurea Foxglove
Ephedrine Bronchodilator Ephedra sinica Ephedra plant
Ipecac Emetic Cephaelis ipecachuanha Ipecac plant
Morphine Analgesic Papaver somniferum Opium poppy
Pseudoephedrine Decongestant Ephedra sinica Ephedra plant
Quinine Antimalarial prophylactic Cinchona pubescens Chinchona
Reserpine Treats hypertension Rauwolfia serpentina Rauwolfia
Sennoside A, B Laxative Cassia angustifolia Senna
Scopolamine Treats motion sickness Datura stramonium Thorn apple
THC Antiemetic Cannabis sativa Marijuana
Toxiferine Relaxes muscles during Strychnos guianensis Strychnos plant
surgery
Tubocurarine Muscle relaxant Chondrodendron tomentosum Curare
Vincristine Treats pediatric leukemia Catharanthus roseus Rose periwinkle
Warfarin Anticoagulant Melilotus spp. Sweet clover
Sources: Balick and Cox 1996; Chivian and Bernstein 2008.
The Value of Biodiversity 67

source of two potent drugs that have increased the rate of survival of child-
hood leukemia from 10% to 90%. Venomous animals such as rattlesnakes,
bees, and cone snails have been especially rich sources of chemicals with
valuable medical and biological applications. An enzyme derived from a
heat-tolerant bacterium (Thermus aquaticus) collected from hot springs at
Yellowstone National Park forms a key component in the polymerase chain
reaction used to amplify DNA in the biotechnology industry and in biologi-
cal research (Figure 3.6). This enzyme is also used in the medical field to
detect human diseases. The industries using this enzyme have generated
hundreds of billions of dollars of value and employ hundreds of thousands
of people. How many more such valuable species will be discovered in the
years ahead—and how many will go extinct before they are discovered?

Figure 3.6 Specialized bacteria, such as these bright orange bacteria at


Prism Lake, grow abundantly in Yellowstone National Park’s hot mineral springs
and have contributed essential enzymes for the high temperature reactions used
in the biotechnology industry. (Photograph by Richard Primack.)
68 Chapter 3

Indirect use values


Many organisms provide benefits to humans only when alive and in their
natural ecosystem, such as plants growing on a hillside that prevent soil
erosion. These resources have nonconsumptive use value because they
are not consumed. Indirect use values are nonconsumptive use values
that can be assigned to all aspects of biodiversity that provide economic
benefits without being harvested or destroyed. Because these benefits are
not goods or services in the usual economic sense, they do not typically
appear in the statistics of national economies, such as the GDP. They are
often called public goods because they belong to society in general, without
private ownership. However, these benefits may be crucial to the contin-
ued availability of the natural products on which economies depend. If
natural ecosystems are not available to provide these benefits, substitute
sources must be found—often at great expense—or local and even regional
economies may face a decline in prosperity or even collapse.
Human societies are totally dependent on the free services that we
obtain from natural ecosystems because we could not pay to replace these
ecosystems if they were permanently degraded or destroyed. Especially
important in this regard are wetland ecosystems because of their role in
water purification and nutrient recycling as well as their enormous impor-
tance in flood control (Horwitz and Finlayson 2011; see the section “Water
and soil protection,” p. 70).
Economists are actively improving calculations of the indirect use
value of ecosystem services at regional and global levels (Bateman et al.
2013; Reyers et al. 2013). However, many ecological economists sharply
disagree about how calculations of indirect use value should be done, or
even whether they should be done at all (Peterson et al. 2010). One such
calculation suggests that the annual value of ecosystem services worldwide
is as much as $145 trillion (Costanza et al. 2014), exceeding the current $78
trillion annual value of the world’s economy (data.worldbank.org). Using
different approaches, other ecological economists have come up with much
lower estimates, but those estimates have still amounted to trillions of dol-
lars a year. The disparity in these various estimates indicates that much
more work needs to be done on this topic.
The great variety of environmental services that ecosystems provide
can be assigned particular types of indirect use value. The following sec-
tions discuss some of the specific indirect use values derived from biodi-
versity. Later in the chapter, we will consider two other nonconsumptive
ways of valuing biodiversity: option value, the value that biodiversity may
have in the future, and existence value, the amount that people are willing
to pay to protect biodiversity (or other environmental goods or services)
even if they never expect to experience it.

Ecosystem productivity
All life on Earth is made possible by the energy of the sun, which is con-
verted into usable energy through photosynthesis in plants and algae. Hu-
The Value of Biodiversity 69

mans depend on the energy stored in plants for many direct uses, such as
food, fuelwood, and hay and other fodder for animals. This plant material
is also the starting point for innumerable food chains, from which people
harvest many animal products. Humans appropriate approximately half
of the productivity of the terrestrial environment to meet their needs for
natural resources (MEA 2005), and most of the remaining half performs
services that have indirect use value to humans, including the production
of oxygen (O2) by plants through the process of photosynthesis.
The destruction of the vegetation in an area through overgrazing by
domestic animals or overharvesting of timber will destroy the system’s
ability to perform these functions (Figure 3.7). Eventually, it will lead
to losses of plant biodiversity and of the associated production of plant
biomass, loss of the animals that live in that area, and losses of
natural resources and ecosystem services for people.
Likewise, coastal estuaries are areas of rapid plant and algal Ecosystems with reduced
growth that provide the starting point for food chains leading species diversity are
to commercial stocks of fish and shellfish. When these coastal less able to adapt to
areas are filled in for development, their value to society is lost. the altered conditions
Even when degraded or damaged wetland ecosystems are re- associated with rising
built or restored—usually at great expense—they often do not carbon dioxide levels and
function as well as they initially did and almost certainly do not global climate change.
contain their original species composition or species richness.
Scientists are actively investigating how the loss of species
from biological communities affects ecosystem processes such
as the total growth of plants, the ability of plants to absorb atmospheric
CO2, and the ability of communities to adapt to global climate change (King
et al. 2012). Many studies of natural and experimental grassland communi-

Figure 3.7 Although many


grassland systems are adapted
to grazing, overgrazing reduces
ecosystem services and can lead
to decreased ecosystem produc-
tivity and increased soil erosion.
(Photograph by Anna Sher.)
70 Chapter 3

(A) (B)
65

Total plant cover (% of surface area)


60 Plant cover approaches
a maximum value
55

50

45
More species, increased plant cover
40

35 Few species and low plant cover


30
25
0 5 10 15 20 25
Number of species present

Figure 3.8 (A) Healthy prairie ecosystems are


naturally diverse, with many species of grasses and
forbs. (B) Varying numbers of grassland species were
grown in experimental plots. The plots containing
the most species had the greatest overall amount of
growth as measured by the total plant cover (the per-
centage of the total surface area occupied by plants).
(A, photograph © Clint Farlinger/Alamy Stock Photo;
B, after Tillman 1999.)

ties confirm that as species are lost, overall productivity declines, and the
community is less flexible in responding to environmental disturbances
such as drought (Hautier et al. 2015) (Figure 3.8).

Water and soil protection


Biological communities are of vital importance in protecting watersheds,
buffering ecosystems against extremes of flood and drought, and maintain-
ing water quality (Thorp et al. 2010). Plant foliage and dead leaves intercept
the rain and reduce its impact on the soil. Plant roots and soil organisms
aerate the soil, increasing its capacity to absorb water. This increased water-
holding capacity reduces the flooding that would otherwise occur after
heavy rains and allows a slow release of water for days and weeks after
the rains have ceased.
When logging, farming, and other human activities disturb vegeta-
tion, the rates of soil erosion, and even occurrences of landslides, increase
rapidly, decreasing the value of the land for some human activities. Dam-
age to the soil limits the ability of plant life to recover from disturbance
and can render the land useless for agriculture. Erosion and flooding also
contaminate drinking water supplies for humans in the communities along
rivers, leading to an increase in human health problems. Soil erosion in-
The Value of Biodiversity 71

creases sediment loads entering the reservoirs behind dams, causing a loss
of electrical output, and it creates sandbars and islands, which reduces the
navigability of rivers and ports.
Floods are currently the most common natural disaster in the world,
killing thousands of people each year, and losses of wetland and floodplain
ecosystems have contributed to these disasters. In the industrial nations of
the world, wetlands protection has become a priority in order to prevent
flooding of developed areas (Figure 3.9). In certain locations, wetlands
are estimated to have a value of $6000 per hectare per year in flood damage
reduction and other ecosystem services, which is three times the value of
farmland developed on the same site (MEA 2005). The conversion of flood-
plain habitat to farmland along the Mississippi, Missouri, and Red Rivers
in North America and the Rhine River in Europe is considered a major
factor in the massive, damaging floods along those rivers in past years. The
most dramatic example of such flooding is the devastating flooding of New
Orleans in 2005 after Hurricane Katrina struck the Mississippi delta, which

Figure 3.9 Wetlands perform many vital functions for humans. The trees,
their root mats, and other vegetation act like a living sponge that traps and slow-
ly releases rainwater and snowmelt, while creating resistance that slows water
flows, thus preventing damage from floodwater. Aquatic plants also clean water
by taking up excess nitrogen, phosphorus, heavy metals, and other substances
that can be harmful to humans. (© Patricia Hofmeester/Shutterstock.)
72 Chapter 3

has undergone heavy conversion of wetlands for urban, indus-


Wetland ecosystem trial, and agricultural development. The risk of flooding would be
services whose value is substantially reduced if even a small proportion of the wetlands
typically not accounted along these rivers were restored to their original condition.
for in the current Wetlands also perform important functions for filtering
market system include excess nutrients and toxins in water. The government of New
waste treatment, water York City paid $1.5 billion in the late 1980s to county and town
purification, and flood governments in rural New York State to maintain forests on the
control—all of which
watersheds surrounding its reservoirs and to improve agricul-
tural practices in order to protect the city’s water supply. Water
are essential to healthy
filtration plants doing the same job would have cost $8–9 billion
human societies.
(www.nyc.gov/watershed). The nonconsumptive use value of
the US national forests alone for protecting the nation’s water
supply has been estimated at $4 billion per year.
Aquatic ecosystems such as swamps, lakes, rivers, floodplains, tidal
marshes, mangroves, estuaries, the continental shelf, and the open ocean
are capable of breaking down and immobilizing toxic pollutants, such as
the heavy metals and pesticides that have been released into the environ-
ment by human activities (Balmford et al. 2002). When such aquatic com-
munities, especially the bacteria and other microorganisms they contain,
are damaged by a combination of sewage overload and habitat destruction,
alternative systems have to be developed. These contrived systems, such as
waste treatment facilities and giant landfills, cost tens of billions of dollars.
In regions that cannot afford to build such facilities, people’s quality of life
can be severely harmed.

Climate regulation
Plant communities are important in moderating local, regional, and even
global climate conditions (West et al. 2011). At the local level, trees provide
shade and evaporate water from their leaf surfaces during photosynthesis,
reducing the local temperature in hot weather. This cooling effect reduces
the need for fans and air conditioners and increases people’s comfort and
work efficiency. Trees are also locally important because they act as wind-
breaks for agricultural fields, reducing soil erosion by wind and reducing
heat loss from buildings in cold weather.
At the regional level, plants capture water that falls as rain and then
transpire it back into the atmosphere, from which it can fall as rain again.
The loss of vegetation from large forested regions such as the Amazon
basin and western Africa may result in a reduction of average annual
rainfall or greatly altered weather patterns over large areas.
In both terrestrial and aquatic environments, plant growth is tied to the
carbon cycle. A reduction in plant life results in reduced uptake of CO2,
contributing to the rising CO2 levels that lead to global warming (McKinley
et al. 2011; Pan et al. 2011). Environmental economists also recognize the
value of intact and restored forests in retaining carbon and absorbing atmo-
spheric CO2 (Butler et al. 2009). As countries and corporations reduce their
The Value of Biodiversity 73

CO2 emissions as part of the worldwide effort to address global climate


change, they are paying to protect and restore forests and other ecosystems
(Venter et al. 2010).

Species relationships and environmental monitors


Many of the species harvested by people for their direct con- Relationships between
sumptive use value depend on other wild species for their con- species are often
tinued existence. For example, the wild game and fish harvested essential for preserving
by people are dependent on wild insects and plants for their biodiversity and
food. Thus, a decline in a wild species of little immediate value providing value to people.
to humans may result in a corresponding decline in a harvested For example, many
species that is economically important. insects pollinate the
Crop plants benefit from wild insects, birds, and bats (Kross crops on which people
et al. 2012; Wanger et al. 2014). Predatory insects such as praying depend for food.
mantises, as well as many bird and bat species, feed on pest
insect species that attack crops, increasing crop
yields and reducing the need to spray pesticides
(Boyles et al. 2011; Kross et al. 2012). Insects, birds, (A)
and bats also act as pollinators for numerous crop
species (Figure 3.10). About 150 species of crop
plants in the United States require insect pollina-
tion of their flowers, which is often performed by
a combination of wild insects and domesticated
honeybees (Garibaldi et al. 2013; Vanbergen and
the Insect Pollinator Initiative 2013). The global
value of these pollinators in increasing crop yield
has been estimated at $200 billion per year. The
value of wild insect pollinators will increase in
the near future if they take over the pollination
role of honeybees, whose populations are declin-
ing in many places because of disease and pests.
Many useful wild plant species depend on fruit- (B)
30

Figure 3.10 (A) Bees provide ecosystem ser-


25
vices for humans by pollinating food crops such as
this peach tree; without pollination, the tree will not
Fruit set (%)

produce fruit. The indirect economic value of these 20


insects can be estimated by what it would cost to
pay humans to hand-pollinate the trees—something
15
that is sometimes necessary when bee populations
decline. (B) Higher visitation rates by wild bees (the
number of bees observed in 15-minute intervals for- 10
aging on 1000 flowers) increase the fruit set of sweet
cherries (the percentage of flowers that develop into 0
fruits). (A, photograph by David McIntyre; B, after 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Holzschuh et al. 2012.) Wild bee visitation rate
74 Chapter 3

eating birds, primates, and other animals to act as seed dispersers. Where
these animals have been overharvested, fruits remain uneaten, seeds are
not dispersed, and plant species head toward local extinction (Sethi and
Howe 2009). It should be noted, however, that there is redundancy in guilds
of similar species, and the service of one natural predator, pollinator, or
seed disperser may be carried out equally well by another species.
One of the most economically significant relationships in ecosystems is
the one between many forest trees and crop plants and the soil organisms
that provide them with essential nutrients (Beattie and Ehrlich 2010). Fungi
and bacteria break down dead plant and animal matter in the soil, which
they use as their energy source. In the process, the fungi and bacteria
release mineral nutrients such as nitrogen into the soil. These nutrients are
used by plants for further growth. The poor growth and dieback of many
trees observed in certain areas of North America and Europe is attributable
in part to the deleterious effects of acid rain and air pollution on soil fungi
that help supply the trees with mineral nutrients and water.

Amenity value
Ecosystems provide many recreational services for humans; for instance,
they furnish a place to enjoy nonconsumptive activities such as hiking,
photography, and birdwatching (Buckley 2009) (Figure 3.11). This expe-

Figure 3.11 A picnic area inside a national park in Turkey. It is argued that
humans have a need to be near natural features such as this lake and forest;
their use for recreation gives these natural features amenity value, a type of indi-
rect use value. (Photograph by Richard Primack.)
The Value of Biodiversity 75

rience of nature is not only enjoyable, but also leads to improved health
for the participants (Donovan et al. 2013). The monetary value of these
activities, sometimes called their amenity value, can be considerable and
can have a major impact on local economies. In the United States, more
than 250 million people visit national parks each year. People in the United
States spend around 7 billion hours per year enjoying nature at national
parks, state parks, wildlife refuges, and other protected public lands (Siika-
mäki 2011). If we estimate that these nature experiences have a value of
$12 per hour (the amount people might spend at a movie or dinner), then
US protected areas have an estimated value of $84 billion per year for just
this one type of use! Recreation represents over 75% of the value of US
national forests, far greater than the value of the wood being extracted
(Groom et al. 2006). Even sportfishing and hunting, which in theory are
consumptive uses, are in practice both consumptive and nonconsumptive
because the food value of the animals caught by fishermen and hunters
is insignificant compared with the time and money spent on these activi-
ties. In national and international sites known for their conservation value
or exceptional scenic beauty, such as Yellowstone National Park, noncon-
sumptive recreational value often dwarfs the value generated or captured
by all other economic enterprises there, including ranching, mining, and
logging (Power and Barrett 2001).
Ecotourism is a special category of recreation that involves people visit-
ing places and spending money wholly or in part to experience unusual
biological communities (such as rain forests, African savannas, coral reefs,
deserts, the Galápagos Islands, or the Everglades) and to view particular
“flagship” species (such as elephants on safari trips; Balmford et al. 2009;
www.ecotourism.org) (Figure 3.12). Tourism, valued at $600 billion dollars
per year, is among the world’s largest industries (on the scale of the petro-
leum and motor vehicle industries), and ecotourism currently represents
about 20% of the tourism industry.
Ecotourism has traditionally been a key industry in East African coun-
tries such as Kenya and Tanzania, and it has also become important in
Latin America, including Costa Rica and Belize, and many other parts of
the world. Tourism associated with the Great Barrier Reef in Australia is
estimated to be worth $5.5 billion per year and employs more than 50,000
people, which is 36 times more than the commercial fishing industry in
Australia (Catlin et al. 2013). In addition to international tourism, the rap-
idly growing middle classes in developing countries, such as China and
India, are increasingly traveling within their own countries to visit national
parks and nature reserves (Karanth and DeFries 2011).
The revenue provided by ecotourism has the potential to provide one of
the most immediate justifications for protecting biodiversity, particularly
when ecotourism activities are integrated into overall management plans
(Vianna et al. 2012). In integrated conservation and development projects
(ICDPs), local communities develop accommodations, expertise in nature
guiding, local handicraft outlets, and other sources of income; the income
76 Chapter 3

(A)

(B) Educates the


public in
Tourism programs, linked to conservation
transportation, accommodations,
education, and protected areas,
often with government support
Protects the
environment

Enhances local sense of place


Provides satisfying tourist experiences A sustainable
tourism system
Is marketed effectively
Is operated safely for all
Employs and trains
Receives the
a local workforce
support of the
Is seen to be benefiting the local community
community (schools, clinics, etc.)

Makes money for investors

Figure 3.12 (A) Ecotourism can provide an economic justification for pro-
tecting biodiversity and can also provide benefits to people living nearby. (B)
The diagram illustrates some of the main elements in a successful ecotourism
program. (A, photograph © Andrew Parkinson/Corbis; B, after Braithwaite 2001.)
The Value of Biodiversity 77

from ecotourism allows the local people to give up unsustain-


able or destructive hunting, fishing, or grazing practices (see The rapidly developing
Chapter 9). The local community benefits from the learning of ecotourism industry can
new skills, employment opportunities, greater protection for the provide income to protect
environment, and the development of additional community biodiversity, but possible
infrastructure such as schools, roads, medical clinics, and stores. costs must be weighed
A danger of ecotourism is that the tourists themselves will along with benefits.
unwittingly damage the sites they visit—by trampling wild-
flowers, breaking coral, or disrupting nesting bird colonies, for
instance—thereby contributing to the degradation and distur-
bance of sensitive areas (Nash 2009; Schlacher and Thompson 2012; Shutt et
al. 2014). Tourists might also indirectly damage sites by creating a demand
for fuelwood for heating and cooking, thus contributing to deforestation.
In addition, the presence, affluence, and demands of tourists can transform
traditional human societies in tourist areas by changing employment op-
portunities (Dahles 2005). As local people increasingly enter a cash-based
economy, their values, customs, and relationship to nature may be lost
along the way. A final potential danger of this industry is that ecotourist
facilities may provide a sanitized fantasy experience rather than helping
visitors understand the serious social and environmental problems that
endanger biodiversity.

Educational and scientific value


Many books, television programs, movies, and websites produced for edu-
cational and entertainment purposes are based on nature themes (Oster-
lind 2005). These natural history materials are continually incorporated
into school curricula and are worth billions of dollars per year. To take one
example, recent movies with penguins as main characters or themes have
had revenues estimated at around $1.6 billion. They represent a noncon-
sumptive use value of biodiversity because they use nature only as intel-
lectual content. A considerable number of professional scientists, as well
as highly motivated amateurs, are engaged in making ecological observa-
tions and preparing educational materials. In rural areas, their activities
often take place in scientific field stations, which can become sources of
training and employment for local people. While these scientific activities
provide economic benefits to the communities surrounding field stations,
their real value lies in their ability to increase human knowledge, enhance
education, and enrich the human experience.

Multiple uses of a single resource: A case study


Horseshoe crabs (Limulus polyphemus) provide an example of the diverse
values that can be provided by just one species. They are usually noticed
only as clumsy creatures that seem to move with difficulty in shallow
seawater (Figure 3.13). In the United States, commercial fishermen har-
vest these animals in large quantities for use as cheap fishing bait. In
recent years, however, ecologists have realized that horseshoe crab eggs
78 Chapter 3

and juveniles are extremely important as a food source for


shorebirds and coastal fish, which have a major role in local
tourism related to birdwatching and sportfishing. Without
horseshoe crabs, shorebird and sport fish populations de-
cline in abundance (Niles 2009). Additionally, the blood of
horseshoe crabs is collected to make limulus amoebocyte
lysate (LAL), a highly valuable chemical used to detect bac-
terial contamination in injection-administered medications
and vaccines (Odell et al. 2005). This chemical cannot be
manufactured synthetically, and horseshoe crabs are its
only source. Without this natural source of LAL, our abil-
ity to determine the purity of injected medicines would be
compromised.
Currently, commercial fishing and sportfishing inter-
ests, environmental groups, birdwatching groups, and the
biomedical industry are competing for control of horseshoe
crabs along US coastlines. Each group can make a good ar-
gument for its own right to use or protect horseshoe crabs.
Hopefully, the final result will be a working compromise
Figure 3.13 Horseshoe crabs that allows the crabs a place in a functioning ecosystem and
gather in great numbers to spawn still provides for the needs of people living in the area and
in shallow coastal waters. These ag- elsewhere in the world.
gregations have significant value to
people. (© Prisma Bildagentur AG/
Alamy.) The Long-Term View: Option Value
The potential of biodiversity to provide an economic benefit
to human society at some point in the future is its option
value. As the needs of society change, so must the methods
of satisfying those needs, and such methods often lie in pre-
Sometimes a specific
viously untapped animal or plant species. For example, the
species or ecosystem can
continued genetic improvement of cultivated plants is neces-
provide a variety of goods
sary not only for increased yield, but also to guard against
and services to human
pesticide-resistant insects and more virulent strains of fungi,
society. Compromises are viruses, and bacteria (Sairam et al. 2005).
often needed to balance We are continually searching the biological communi-
competing uses. ties of the world for new plants, animals, fungi, and micro-
organisms that can be used to fight human diseases or to
provide some other economic value, an activity referred to
as bioprospecting (Lawrence et al. 2010). These searches are
generally carried out by government research institutes and pharmaceuti-
cal companies. The US National Cancer Institute has been carrying out a
program to test extracts of thousands of wild species for their effective-
ness in controlling cancer cells and HIV (the virus that causes AIDS). To
facilitate the search for new medicines and to profit financially from new
products, GlaxoSmithKline, a British pharmaceutical corporation, and the
Brazilian government signed a contract in 1999 worth $3 million allow-
The Value of Biodiversity 79

ing the company to sample, screen, and investigate approximately 40,000


plants, fungi, and bacteria from Brazil, with part of the royalties going to
support scientific research and local community-based conservation and
development projects. Similar agreements involving international corpora-
tions have been signed in many other countries as well. Another approach
has been to target plants and other natural products used in traditional
medicine for screening, often in collaboration with local healers and rural
villagers. Programs such as these provide financial incentives for countries
to protect their natural resources and the biodiversity knowledge of their
indigenous inhabitants.
The search for valuable natural products is wide-ranging:
entomologists search for insects that can be used as biological A question currently
control agents, microbiologists search for bacteria that can assist being debated among
in biochemical manufacturing processes, and wildlife biologists conservation biologists,
search for species that can potentially produce animal protein governments, ecological
more efficiently and with less environmental damage than ex- economists, corporations,
isting domesticated species (Chivian and Bernstein 2008). The and local individuals is,
growing biotechnology industry is finding new ways to reduce Who owns the commercial
pollution, to develop better industrial processes, and to fight dis- rights to the world’s
eases threatening human health. The gene-splicing techniques biodiversity?
of molecular biology are allowing unique, valuable genes found
in one species to be transferred to another species. Both newly
discovered and well-known species are often found to have exactly those
properties needed to address some significant human problem. If biodiver-
sity is reduced, the ability of scientists to locate and utilize a broad range
of species will also be reduced.
A question currently being debated among conservation biologists,
governments, environmental economists, and corporations is, Who owns
the commercial development rights to the world’s biodiversity? The answer
to this is complex, especially in the modern era. An excellent example
is provided by the immunosuppressant drug cyclosporine, which occurs
naturally in the fungus Tolypocladium inflatum. The Swiss company Sandoz
(which later merged with another company to become Novartis) developed
cyclosporine into a family of drugs with sales of $1.2 billion per year (Bull
2004). Cyclosporine is a key medicine used following transplant surgery
to prevent organ rejection. The fungus was found in a sample of soil that
was collected in Norway, without permission, by a Sandoz biologist on
vacation and was later screened for biological activity. Norway has not yet
received any payment for the use of this fungus in drug production. Simi-
larly, the US National Park Service has not received any payments for the
bacteria collected at Yellowstone National Park that led to breakthroughs
in DNA technology. Such past and present unauthorized bioprospecting
for commercial purposes is now often termed biopiracy. Many developing
countries have reacted to this situation by passing laws that require permits
for collecting biological material for research and commercial purposes,
and they impose criminal penalties and fines for the violation of such laws
80 Chapter 3

(Bhatti et al. 2009). People collecting samples without the needed permits
have been arrested for violating the law.
Both developing and developed countries now frequently demand a
share in the commercial activities derived from the biodiversity contained
within their borders, and rightly so. Local people in developing countries
who possess knowledge of species, protect them, and show them to scien-
tists should also share in the profits from any use of them. Writing treaties
and developing procedures to guarantee participation in this process will
be a major diplomatic challenge in the coming years.
While most species may have little or no direct economic value and
little option value, a small proportion may have enormous potential to
supply medical treatments, to support a new industry, or to prevent the
collapse of a major agricultural crop. Other species or sets of species may
provide other kinds of future values even if they don’t provide them now.
If just one of these species goes extinct before it is discovered, it could be a
tremendous loss to the global economy, even if the majority of the world’s
species are preserved. As Aldo Leopold commented in Round River,
If the biota, in the course of aeons, has built something we like but do
not understand, then who but a fool would discard seemingly useless
parts? To keep every cog and wheel is the first precaution of intelligent
tinkering.
The diversity of the world’s species can be compared to a manual on
how to keep the Earth running effectively. The loss of a species is like
tearing a page out of the manual. If we ever need the information from
that page to save ourselves and the Earth’s other species, the information
will have been irretrievably lost.

Existence Value
Many people throughout the world care about wildlife, plants, and entire
ecosystems and want to see them protected. Their concern may be asso-
ciated with a desire to someday visit the habitat of a unique species and
see it in the wild; alternatively, concerned individuals may not expect,
need, or even desire to see a species personally or experience
the habitat in which it lives. For this reason, existence value is
considered a non-use value: people value the resource with-
People, governments, and out any intention to use it now or in the future. In economic
organizations annually terms, existence value is the amount that people are willing to
contribute large sums pay to prevent species from going extinct, habitats from being
of money to ensure the destroyed, and genetic variation from being lost (Zander and
continuing existence Garnett 2011). A related idea is beneficiary value, or bequest
of certain species and value: how much people are willing to pay to protect some-
ecosystems. thing of value for their own children and descendants, or for
future generations.
Particular species—the so-called charismatic megafauna,
such as pandas, whales, lions, and many birds—elicit strong
The Value of Biodiversity 81

Figure 3.14 Whale watching can make a significant contribution to a local


economy, and participants often later contribute money to organizations promot-
ing whale conservation. Here, people greet a California gray whale (Eschrichtius
robustus) in Magdalena Bay, on the Pacific side of the Baja Peninsula. Such meet-
ings can enrich human lives. (© Robert Harding Specialist Stock/Corbis.)

responses in people (Figure 3.14). Special groups have been formed to


appreciate and protect butterflies and other insects, wildflowers, and fungi.
People place value on wildlife in a direct way by joining and contributing
billions of dollars to organizations that protect species (Wallmo and Lew
2012). In the United States, billions of dollars are contributed each year to
conservation and environmental organizations, with The Nature Conser-
vancy ($871 million in 2013), the World Wildlife Fund ($245 million), Ducks
Unlimited ($184 million), and the Sierra Club ($98 million) high on the list.
Citizens also show their concern by directing their governments to spend
money on conservation programs and to purchase land for habitat. For
example, the government of the United States has spent millions of dollars
to protect a single rare species, the brown pelican (Pelecanus occidentalis),
which was protected initially under the US Endangered Species Act and
is now considered recovered.
Existence value can also be attached to ecosystems, such as temperate
old-growth forests, tropical rain forests, coral reefs, and prairie remnants,
and to areas of scenic beauty. Growing numbers of people and organiza-
tions contribute large sums of money annually to ensure the continuing
existence of these habitats, and the environment is increasingly an impor-
tant issue during national elections. Further, people want environmental
education included in public school curricula (www.neefusa.org; www.
epa.gov/education).
82 Chapter 3

In summary, ecological economics has helped to draw attention to the


wide range of goods and services provided by biodiversity. That attention
has enabled scientists to account for environmental impacts that were pre-
viously left out of the equation. When analyses of large-scale development
projects have finally been completed, some projects that initially appeared
to be successful have been seen to actually be running at an economic
loss. For example, to evaluate the success of an irrigation project using
water diverted from a tropical wetland ecosystem, the short-term benefits
(improved crop yields) must be weighed against the environmental costs.
Figure 3.15 shows the total economic value of a tropical wetland ecosys-
tem, including its use value, option value, and existence value. When the
wetland ecosystem is damaged by the removal of water, the ecosystem’s
ability to provide the goods and services shown in the figure are curtailed,

Total Economic Value of a Tropical Wetland Ecosystem

Use Values Existence Value


Direct Use Values Indirect Use Values Protection of biological
diversity
Fish and meat Flood control
Maintaining culture of
Fuelwood Soil fertility local people
Option Value
Timber and other Pollution control Continuing ecological and
building materials Future products:
Drinking water evolutionary processes
Medicinal plants Medicines
Transportation
Edible wild fruits Genetic resources
Recreation and tourism
and plants Biological insights
(e.g., birdwatching)
Animal fodder Food sources
Education
Biological services Building supplies
(e.g., pest control, pollination) Water supplies

Figure 3.15 Evaluation of the success of a development project must in-


corporate the full range of its environmental impacts. This figure shows the total
economic value of a tropical wetland ecosystem, including direct and indirect
use value, option value, and existence value. A development project such as an
irrigation project lowers the value of the wetland ecosystem when water is re-
moved for crop irrigation. When that lowered value is taken into account, the ir-
rigation project may represent an economic loss. (After Groom et al. 2006; based
on data in Emerton 1999.)
The Value of Biodiversity 83

their value greatly diminishes, and the economic success of the project is
called into question. It is only by incorporating the value of the wetland
into this equation that an accurate view of the total project can be gained.

Environmental Ethics
In most modern societies, people attempt to protect biodiversity, environ-
mental quality, and human well-being through regulations, incentives,
fines, environmental monitoring, and assessments. A complementary ap-
proach is to change the fundamental values of our materialistic society.
Environmental ethics, a vigorous and growing discipline within philoso-
phy, articulates the ethical value of the natural world (Alexander 2009;
Minteer and Collins 2008). As a corollary, it challenges the materialistic
values that tend to dominate modern societies. If contemporary societies
de-emphasized the pursuit of wealth and instead focused on furthering
genuine human well-being, the preservation of the natural environment
and the maintenance of biodiversity would probably become honored prac-
tices, rather than occasional afterthoughts (Mills 2003).

Ethical values of biodiversity


Environmental ethics provides virtues and values that make
sense to people today. At a time when there are unprecedent- Ethical arguments
ed threats to the environment, ethical arguments can and do can complement
convince people to conserve biodiversity. Ethical arguments economic and biological
are also important because, although economic arguments by arguments for protecting
themselves provide a basis for valuing some species and eco- biodiversity. Such ethical
systems, economic valuation can also provide grounds for ex- arguments are readily
tinguishing species, or for saving one species and not another understood by many
(Redford and Adams 2009; Rolston 2012). According to conven- people.
tional economic thinking, a species with low population num-
bers, an unattractive appearance, no immediate use to people,
and no relationship to any species of economic importance will
be given a low value. Halting profitable developments or making costly
attempts to preserve these species may not have any obvious economic
justification. In fact, in many circumstances, economic cost–benefit analy-
ses will support the destruction of endangered species that stand in the
way of “progress.”
Despite any economic justification, however, many people would make
a case against species extinctions on ethical grounds, arguing that the
conscious destruction of a natural species is morally wrong, even if it is
economically profitable. Similar arguments can be advanced for protecting
unique ecosystems and genetic variation. Ethical arguments for preserv-
ing biodiversity resonate with people because they appeal to our nobler
instincts or to belief in a divine creation, which do play a role in societal
decision making (Bhagwat et al. 2011).
The following arguments, based on the intrinsic value of species and on
our duties to other people, are important to conservation biology because
84 Chapter 3

they provide a rationale for protecting all species, including rare species
and species of no obvious economic value.

Each species has a right to exist All species represent unique


biological solutions to the problem of survival. For this reason, the sur-
vival of each species must be respected, regardless of its importance to
humans. This statement is true whether the species is large or small,
simple or complex, ancient or recently evolved; whether it is economically
important or of little immediate economic value to humans; and
whether it is loved or hated by humans. Each species has value
for its own sake—an intrinsic value unrelated to human needs
An argument can be or desires (Sagoff 2008). This argument suggests not only that
made that people have a we have no right to destroy any species, but also that we have
responsibility to protect a moral responsibility to actively protect species from going
species and other aspects extinct as the result of our activities, as articulated by the deep
of biodiversity because of ecology movement described later in this section. This argu-
their intrinsic value, not ment also recognizes that humans are part of the larger biotic
because of human needs. community and reminds us that we are not the center of the
universe (Figure 3.16).

WRONG RIGHT

Figure 3.16 According to certain principles of environmental ethics, it is


wrong for humans to act as if they are at the top of the living world and have the
right to exploit and damage other species; rather, there is a moral imperative for
humans to behave as if all species have an equal right to exist. (Silhouettes ©
Shutterstock.)
The Value of Biodiversity 85

All species are interdependent Species interact in com-


plex ways in natural communities. The loss of one species may All species and eco-
have far-reaching consequences for other members of the commu- systems are interdepen-
nity (as described in Chapter 2). Other species may become extinct dent, and so all parts
in response, or the entire ecosystem may become destabilized as of nature should be
the result of cascades of species extinctions. For these reasons, if protected. It is in the
we value some parts of nature, we should protect all of nature long-term survival
(Leopold 1949). We are obligated to conserve the system as a whole
interest of people to
because that is the appropriate survival unit (Diamond 2005).
protect all of biodiversity.
People have a responsibility to act as stewards
of the Earth Many religious adherents find it wrong to de-
stroy species because they are God’s creations (Moseley 2009;
see Chapter 1). If God created the world, then presumably the species God
created have value. Within the Jewish, Islamic, and Christian traditions,
human responsibility for protecting animal species is explicitly described
in the Bible as part of the covenant with God. For example, Muslim clerics
in Malaysia recently joined Indonesia in a fatwa (a religious edict) against
hunting endangered species; Asrorun Niam Sholeh of the Indonesian Ulema
Council stated, “As Muslims, we have a duty to maintain the ecological bal-
ance” (Yi 2016). Other major religions, including Hinduism, Buddhism, and
Islam strongly support the preservation of nonhuman life. As Mahatma
Gandhi (1869–1948) taught, “The good man is the friend of all living things.”

People have a responsibility to future generations If in


our daily living we degrade the natural resources of the Earth and cause
species to become extinct, future generations will pay the price in terms
of a lower standard of living and quality of life (Gardiner et al. 2010). As
species are lost and wild lands developed, children are deprived of one of
the most exciting experiences in growing up: the wonder of seeing “new”
animals and plants in the wild. To remind us to act more responsibly, we
might imagine that we are borrowing the Earth from future generations
who expect to get it back in good condition.

Respect for human life and human diversity is compatible


with a respect for biodiversity Some people worry that recog-
nizing intrinsic value in nature requires taking resources and opportunities
away from human beings. But respect for and protection of biodiversity
can be linked to greater opportunities and better health for people (Jacob
et al. 2009). Some of the most exciting developments in conservation biol-
ogy involve supporting the economic development of disadvantaged rural
people in ways that are linked to the protection of biodiversity. In devel-
oped countries, the environmental justice movement seeks to empower poor
and politically weak people, who are often members of minority groups,
to protect their own environments; in the process, their well-being and the
protection of biodiversity are enhanced (Robinson 2011).
86 Chapter 3

People benefit from aesthetic and recreational enjoy-


ment of biodiversity Throughout history, poets, writers, painters,
and musicians of all cultures have drawn inspiration from wild nature
(Swanson et al. 2008; Thoreau 1854). Nearly everyone enjoys wildlife and
landscapes aesthetically, and this joy increases the quality of our lives.
Nature-related activities are important in childhood development (Carson
1965; Luck et al. 2011) and may even improve human health (Donovan et al.
2013). The presence and abundance of trees in urban settings has frequently
been associated with lower incidence of crime (e.g., Gilstad-Hayden et al.
2015). Recreational activities such as hiking, canoeing, and mountain climb-
ing are physically, intellectually, and emotionally satisfying. People spend
tens of billions of dollars annually in these pursuits, proof enough that
they value them highly. A loss of biodiversity diminishes this experience.
What if there were no more migratory birds or no more meadows filled
with wildflowers and butterflies? Would we still enjoy nature as much?

People benefit from the knowledge the natural world


provides Three of the central mysteries in the world of science are:
(1) how life originated, (2) how the diversity of life interacts to form com-
plex ecosystems, and (3) how humans evolved. Thousands of biologists are
working on these questions and are coming ever closer to the answers. New
techniques of molecular biology allow greater insight into the relationships
of living species as well as some extinct species, which are known to us
only from fossils. When species become extinct and ecosystems are dam-
aged, however, important clues are lost, and the mysteries become harder to
solve. For example, if Homo sapiens’ closest living relatives—chimpanzees,
bonobos, gorillas, and orangutans—disappear from the wild, we will lose
important clues to human physical and social evolution.

Deep ecology
One well-developed environmental philosophy that supports environ-
mental activism is known as deep ecology (Naess 2008). Deep ecology
builds on the basic premise of biocentric equality, which expresses “the
intuition…that all things in the biosphere have an equal right
to live and blossom and to reach their own individual forms of
unfolding” (Devall and Sessions 1985). Humans have a right to
Deep ecology is an
live and thrive, as do the other organisms with whom we share
environmental philosophy the planet (see Figure 3.16). Deep ecologists oppose what they
that advocates placing see as the dominant worldview, which places human concerns
greater value on above all and views human happiness in materialistic terms.
protecting biodiversity Deep ecologists see acceptance of the intrinsic value of nature
through changes in less as a limitation than as an opportunity to live better lives.
personal attitude, Paul Shepard (1925–1996) introduced to the deep ecology
lifestyle, and even movement the idea that we should achieve this ideal by return-
societies. ing to a more primitive state; that civilization has made us im-
mature and out of sync with our environment. He argued that
The Value of Biodiversity 87

because we evolved in close contact with nature, this contact is necessary


for our emotional and psychological well-being. A related idea, called “na-
ture deficit disorder,” was popularized by author Richard Louv to describe
the wide variety of problems people suffer when deprived of interactions
with the natural world (Louv 2005).
Because present human activities are destroying the Earth’s biodi-
versity, our existing political, economic, technological, and ideological
structures must change. The changes that deep ecology calls for entail
enhancing the quality of life for all people, emphasizing improvements
in environmental quality, aesthetics, culture, and spirituality rather than
higher levels of material consumption. The philosophy of deep ecology
includes an obligation to work to implement needed programs through
political activism and a commitment to personal lifestyle changes, in the
process transforming the institutions in which we work, study, pray, and
shop (Bearzi 2009). Deep ecology and other ethical and religious perspec-
tives urge professional biologists, ecologists, and all concerned people (such
as you?) to escape from their narrow, everyday concerns and act and live
“as if nature mattered” (Naess 1989).

Summary
„„Ecological economics is developing provides value to recreation, education,
methods for valuing biodiversity and, and ecotourism activities.
in the process, providing arguments „„The option value of biodiversity is its
for its protection. Direct use values are potential to provide future benefits to
assigned to products harvested from human society, such as new medicines,
the wild, such as timber, fuelwood, industrial products, and crops. Biodiver-
fish, wild animals, edible plants, and sity also has existence value, which is
medicinal plants. Direct use values can the amount of money people and their
be further divided into consumptive governments are willing to pay to pro-
use values, for products that are used tect species and ecosystems without any
locally, and productive use values, for plans for their direct or indirect use.
products harvested in the wild and later
„„Environmental ethics appeals to reli-
sold in markets.
gious and secular value systems to justi-
„„Indirect use values can be assigned to fy preserving biodiversity. The most cen-
aspects of biodiversity that provide tral ethical argument asserts that people
economic benefits to people but are not must protect species and other aspects
harvested during their use. Noncon- of biodiversity because they have in-
sumptive use values include ecosys- trinsic value, unrelated to human needs.
tem productivity, protection of soil and Further, biodiversity must be protected
water resources, positive interactions of because human well-being is linked to a
wild species with commercial crops, and healthy and intact environment.
regulation of climate. Biodiversity also
88 Chapter 3

For Discussion
1. Find a recent large development project ecosystem services such as flood con-
in your area, such as a dam, office park, trol, freshwater provisioning, and soil
shopping mall, highway, or housing de- retention.
velopment, and learn all you can about 3. Imagine that the only known population
it. Estimate the costs and benefits of this of a dragonfly species will be destroyed
project in terms of biological diversity, unless money can be raised to purchase
economic prosperity, and human health. the pond where it lives and the sur-
Who pays the costs and who receives the rounding land. How could you assign a
benefits? Consider other projects carried monetary value to this species?
out in the past and determine their im- 4. Do living creatures, species, biological
pact on the surrounding ecosystem and communities, and physical entities, such
human community. as rivers, lakes, and mountains, have
2. Consider the natural resources that rights? Can we treat them any way we
people use near where you live. Can you please? Where should we draw the line
place an economic value on those re- of moral responsibility?
sources? If you can’t think of any prod-
ucts harvested directly, consider basic

Suggested Readings
Bateman, I. J., A. R. Harwood, G. M. Mace, R. T. Watson, D. J. Abson, B. An-
drews, and 19 others. 2013. Bringing ecosystem services into economic
decision-making: land use in the United Kingdom. Science 341: 45–50.
Braunisch, V., J. Coppes, S. Bachle, and R. Suchant. 2015. Underpinning the
precautionary principle with evidence: A spatial concept for guiding wind
power development in endangered species’ habitats. Journal for Nature
Conservation 24: 31–40.
Cannon, J. R. 1996. Whooping crane recovery: a case study in public and pri-
vate cooperation in the conservation of endangered species. Conservation
Biology 10: 813–821.
Chan, K. M., P. Balvanera, K. Benessaiah, M. Chapman, et al. 2016. Opinion:
Why protect nature? Rethinking values and the environment. Proceedings
of the National Academy of Sciences 113: 1462–1465. Considering the value
of nature in terms of relational values (instrumental) as distinct from those
that are of the object itself (intrinsic).
Costanza, R. R. de Groot, P. Sutton, S. van der Ploeg, and 4 others. 2014.
Changes in the global value of ecosystem services. Global Environmental
Change 26: 152–158. Global ecosystem services are valued between $125
trillion and $145 trillion per year, with significant modern losses due to
land conversion and other factors.
Ehrlich, P. R. and A. H. Ehrlich. 1982. Extinction: The Causes and Consequences of
the Disappearance of Species. Gollancz, London.
Foster, K. R., P. Vecchia, and M. H. Repacholi. 2000. Science and the precau-
tionary principle. Science 288: 979–981. Despite its popular use in govern-
ment policies and international agreements, the precautionary principle is
problematic due to its variable interpretation.
The Value of Biodiversity 89

Hardin, G. 1968. The tragedy of the commons. Science 162: 1243–1248. This
often-cited work suggests that population control is the only solution to
the overuse of common-property resources.
Helm, D. 2015. Natural Capital: Valuing the Planet. Yale University Press. 296 pp.
This author attempts to quantify the value of various environmental re-
sources and argues from an economist’s perspective that the environment
should be at the center of any economy, rather than an afterthought.
Merckx, T. and H. M. Pereira. 2015. Reshaping agri-environmental subsidies:
From marginal farming to large-scale rewilding. Basic and Applied Ecology
16: 95–103. Re-purposing European Union subsidies that make it profitable
to farm marginal land will benefit biodiversity.
Threats to
4 Biodiversity
Human Population Growth and Global Climate Change 118
Its Impact 92 Overexploitation 126
Habitat Destruction 96 Invasive Species 132
Habitat Fragmentation 106 Disease 142
Environmental Degradation and A Concluding Remark 146
Pollution 112

Suburban development, like that seen here outside of


Las Vegas, decreases and degrades native habitat.
M
aintaining a healthy environment means
preserving all of its components in good
condition—ecosystems, biological commu-
nities, species, populations, and genetic variation.
If species, ecosystems, and populations are adapted
to local environmental conditions, why are they be-
ing lost? Why don’t they tend to persist in the same
place over time? Why can’t they adapt to a changing
environment? These questions have a single, simple
answer: massive anthropogenic disturbances (that
is, disturbances caused by human activities) have
altered, degraded, and destroyed the landscape on a
vast scale, destroying populations, species, and even
whole ecosystems.
There are seven major threats to biodiversity: habi-
tat destruction, habitat fragmentation, habitat degra-
dation (including pollution), global climate change, the
overexploitation of species for human use, the inva-
sion of nonnative species, and the spread of disease
(Figure 4.1). Most threatened species face at least two
of these threats, which may interact to speed their way
toward extinction and hinder efforts to protect them
(Forister et al. 2010). Typically, these threats develop so
rapidly, and on such a large scale, that species are not
able to adapt genetically to the changes or disperse to
a more hospitable location. Moreover, multiple threats
92 Chapter 4

Increasing human population and consumption


Agriculture Logging Fisheries Industry Urbanization and International trade
and fossil road construction
fuel use

Habitat loss
Invasive species
Habitat fragmentation
Overexploitation Disease
Habitat degradation
(including pollution)

Climate change

Loss of biodiversity
Extinction of species and populations
Degradation of ecosystems
Erosion of genetic diversity and evolutionary potential
Loss of ecosystem services
Erosion of support systems for human societies

FIGURE 4.1 The major threats to biodiversity (yellow boxes) are the result of
human activities. These seven factors can interact synergistically to speed up
the loss of biodiversity. (After Groom et al. 2006.)

may interact additively, or even synergistically, such that their combined


impact on a species or an ecosystem is greater than their individual effects.

Human Population Growth


and Its Impact
The major threats to
biodiversity—habitat The seven major threats to biodiversity are all caused by the
destruction, habitat ever-increasing use of the world’s natural resources by an ex-
fragmentation, pollution, panding human population. Up until the last 300 years, the rate
global climate change,
of human population growth had been relatively slow, with the
birthrate only slightly exceeding the mortality rate. The great-
overexploitation of
est destruction of ecosystems has occurred since the Industrial
resources, invasive
Revolution, during which time the human population exploded
species, and the spread
from 1 billion in 1850 to 7.25 billion in 2016 (see Figure 1.1). The
of disease—are all rooted
human population
Introduction to Conservation Biology 1E could reach a maximum of 9–10 billion by the
Primack/Sher
in the expanding human end of the twenty-first century. Humans have increased in such
Sinauer Associates
population. Morales Studio/SA
numbers because birthrates have remained high while mortality
Primack_Sher1E_04.01 Date 12-17-15 03-28-16
rates have declined—a result of both modern medical achieve-
Threats to Biodiversity 93

TABLE 4.1 Three Ways Humans Dominate the Global


Ecosystem
1. Land surface
Human land use, mainly for agriculture, and our need for resources,
especially forest products, have transformed as much as half of the Earth’s
ice-free land surface.
2. Nitrogen cycle
Each year, human activities, such as cultivating nitrogen-fixing crops, using
nitrogen fertilizers, and burning fossil fuels, release more nitrogen into
terrestrial systems than do natural biological and physical processes.
3. A
 tmospheric carbon cycle
By the middle of the twenty-first century, human use of fossil fuels and
the cutting down of forests will result in a doubling of the concentration of
carbon dioxide in the Earth’s atmosphere.
Sources: Data from MEA 2005 and Kulkarni et al. 2008.

ments (specifically, the control of disease) and more reliable food supplies.
Population growth has slowed in the industrialized countries of the world,
as well as in some developing countries in Asia and Latin America, but it
is still high in other areas, particularly in tropical Africa. If these countries
implemented immediate and effective programs of population control, the
human population could possibly peak at “only” 9.4 billion in 2050 and
then gradually decline.
Humans and their activities dominate ecosystems worldwide (Table
4.1). People use large amounts of natural resources, such as fuelwood,
timber, wild meat, and wild plants, and convert vast areas of natural habitat
into agricultural and residential lands. Agricultural systems and other
human activities now occupy one-fourth of the Earth’s land surface (Kraus-
mann et al. 2013). All else being equal, more people equals greater human
impact, more land clearing for agriculture, and less biodiversity (Allendorf
and Allendorf 2012; Godfray et al. 2010). For example, nitrogen pollution is
greatest in rivers flowing through landscapes with high human popula-
tion densities, and rates of deforestation are greatest in countries with the
highest rates of human population growth. Therefore, some scientists have
argued strongly that controlling the size of the human population is the
key to protecting biodiversity (Ehrlich et al. 2012; O’Neill et al. 2012; but
see Bradshaw and Brook 2014).
Healthy ecosystems can persist close to areas with high population
densities, even large cities, as long as human activities are regulated by
local customs or government officials. The sacred groves of trees that are
preserved next to villages in Africa, India, and China are examples of
locally managed biological communities. When this regulation breaks
down during war, political unrest, or other periods of social instability,
the result is usually a scramble to collect and sell resources that had been
used sustainably for generations. The higher the human population den-
94 Chapter 4

sity, and the larger the city, the more closely human activities must be
regulated, because the potential for both destruction and conservation is
greater (Gaston 2010). For example, larger cities have been found to produce
proportionally greater carbon dioxide (CO2) emissions than smaller ones
(Fragkias et al. 2013).
People in industrialized countries (and the wealthy minority in devel-
oping countries) consume a disproportionately large share of the world’s
energy, minerals, wood products, and food (Mills Busa 2013), and therefore
have disproportionate effects on the environment. Each year, the United
States, which has 5% of the world’s human population, uses roughly 25%
of the world’s natural resources. And each year, the average US citizen
uses 28 times more energy and 79 times more paper products than does
the average citizen of India (Encyclopedia of the Nations 2009; Randolph and
Masters 2008).
The impact (I) of any human population on the environment is roughly
captured by the formula I = PAT where P is the number of people, A is the
average income, and T is the level of technology (Davidson and Andrews
2013; Elrich and Holdren 1971). It is important to recognize that the impact
of a population is often felt over a great distance; for example, citizens of
Germany, Canada, and Japan affect the environment in other countries
through their use of foods, luxury goods, and other materials produced
elsewhere (Berger et al. 2013). The increasing interconnectedness of re-
source and labor markets is termed globalization. The fish eaten quietly
at home in Washington, DC, may have come from Alaskan waters, where
its capture may have contributed to the population decline of sea lions,
seals, and sea otters; the chocolate cake and coffee consumed at the end
of a meal in Italy or France were made with cacao and coffee
beans that might have grown in plantations carved out of rain
The enormous consump- forests in western Africa, Indonesia, or Brazil. Residents of in-
tion of resources in an dustrialized countries also affect other countries through the
production of waste, including greenhouse gases (with China,
increasingly globalized
first among the top 10 CO2 emitters, producing more than the
world is not sustainable
remaining 9 combined; PBL Netherlands Environmental As-
in the long term.
sessment Agency 2012).
This linkage has been captured in the idea of the ecological
footprint, defined as the per capita influence a group of people
has on both the surrounding environment and locations across the globe
(Holden and Hoyer 2005; Wackernagel and Rees 1996; Figure 4.2). The
ecological footprint per person is high in developed countries such as the
United States and Canada and relatively low in developing countries such
as China and India.
A modern city in a developed country typically has an ecological foot-
print that is hundreds of times its area. For example, the city of Toronto,
Canada, occupies an area of 630 km2, but each of its citizens requires the
environmental services of 7.7 ha (0.077 km 2) to provide food, water, and
waste disposal sites. With a population of 2.4 million people, Toronto has
Threats to Biodiversity 95

12

Central Asia
United Arab Emirates
10 North Africa
Ecological footprint (global hectares per person)

United States
Eastern Europe
East Asia
8 South and Southeast Asia
Latin America Kuwait Canada

Sub-Saharan Africa Australia


6 Middle East
Norway
Other developed countries UK

Saudi Arabia
4 Ireland
Japan
Slovenia
World average South Africa
2
China Argentina
Sierra
Leone
Mozambique Malawi India
0
200 180 160 140 120 100 80 60 40 20 0
Human Development Index rank

FIGURE 4.2 An ecological footprint for a nation is arrived at by calcula-


tions that estimate the number of global hectares needed to support an average
citizen of that nation. Although the methods used to arrive at these calculations
can be debated, the overall message is clear. When plotted against the Hu-
man Development Index, which reflects living standards, ecological footprints
graphically illustrate the disproportionate use of natural resources by people in
developed nations. However, the total impacts (not shown in this graph) of de-
veloping countries such as China, with 1.3 billion people, are also huge because
of their large populations. (Data from Global Footprint Network and the United
Nations Development Programme 2009.)

an ecological footprint of 185,000 km 2, an area equal to the state of New


Jersey or the country of Syria and more than 300 times its own area. This
excessive consumption of resources is not sustainable in the long term.
Unfortunately, this pattern of consumption is now being adopted by the
expanding middle class in the developing world, including the large, rap-
idly developing countries of China and India, increasing the probability
of massive environmental disruption (Feng et al. 2009; Grumbine 2007).
In fact, the developing countries now generate more greenhouse gases
Introduction to Conservation Biology 1E Primack/Sher
and consume
Sinauer Associates more of certain natural resources than the developed coun-
tries; Studio/SA
Morales China, in particular, has emerged as a rapidly growing industrial
Primack_Sher1E_04.02
powerhouse that not only Date 12-17-15manufactured
exports 12-24-15 1/15/16goods but also imports
96 Chapter 4

resources from around the world. The affluent citizens of developed coun-
tries must confront their excessive consumption of resources and reevalu-
ate their lifestyles while at the same time offering aid to curb population
growth, protect biodiversity, and assist industries in the developing world
to grow in a responsible way.
An alternative view is that development can have a positive effect on
biodiversity because wealthier countries can better afford to establish and
maintain their national parks and other natural areas. Across the world,
countries with a gross domestic product (GDP) of $10,000 per capita per
year or greater have stable or increasing forest areas, while countries with
a GDP of less than $10,000 per capita per year have declining forest areas.
Furthermore, developed countries are often characterized by increasing
urbanization and consequent reduced impacts on rural areas.

Habitat Destruction
The primary cause of the reduction in biodiversity, including variation at
the genetic, species, and ecosystem levels, is the habitat loss that inevitably
results from the expansion of human populations and activities (Figure
4.3). For the next few decades, land-use change will continue to be the main
factor affecting biodiversity in terrestrial ecosystems, probably followed
by overexploitation, climate change, and the introduction of invasive spe-
cies (IUCN 2004). Consequently, the most important means of protecting

Mammals Birds Amphibians Gymnosperms

Habitat loss and


degradation
Overexploitation

Invasive species

Disease

Pollution

Intrinsic factors

0 40 80 0 40 80 0 40 80 0 40 80
Percent of threatened species affected

FIGURE 4.3 Habitat loss and degradation are the greatest threats to the
world’s species, followed by overexploitation and intrinsic factors, which include
poor dispersal ability, low reproductive success, and high juvenile mortality.
Groups of species face different threats: birds are more threatened by invasive
species, whereas amphibians are more affected by disease and pollution. Per-
centages add up to more than 100% because many species face multiple threats.
(After IUCN 2004.)
Threats to Biodiversity 97

biodiversity is habitat preservation. Habitat loss does not neces-


sarily mean wholesale habitat destruction—habitat fragmen- The main threat to bio-
tation and habitat damage associated with pollution can also diversity is habitat des-
mean that the habitat is effectively “lost” to species that cannot truction.
tolerate these changes, even though, to the casual onlooker, the
habitat appears intact.
In many areas of the world, particularly on islands and in
locations where the human population density is high, most of the original
habitat has been destroyed (Hambler et al. 2011). Fully 98% of the land
suitable for agriculture has already been transformed by human activity
(Sanderson et al. 2002), and 53.5% of Earth’s surface has been modified
by humans (Hooke et al. 2012). Because the world’s population, as well as
its standard of living, will continue to increase, the world’s farmers will
need to increase agricultural output by 30%–50% over the next 30 years.
Thus, the need to protect biodiversity will be forced to compete directly
against the need for new agricultural lands and the intensification of
agriculture on existing lands.
Habitat disturbance has been particularly severe throughout Europe;
in southern and eastern Asia, including the Philippines, China, and Japan;
in southeastern and southwestern Australia; in New Zealand; in Madagas-
car; in western Africa; on the southeastern and northern coasts of South
America; in Central America; in the Caribbean; and in central and eastern
North America (Figure 4.4). In many of these regions, more than 50%
of the natural habitats have been disturbed or removed. Only 15% of the
land area in Europe remains unmodified by human activities, and in some
regions of Europe, the percentage is even lower. In Germany or the United
Kingdom, for example, one can hardly find any habitat that has not been
modified by humans at one time or another.
The principal human activities that threaten the habitats of endangered
species, in order of decreasing importance, are agriculture (affecting 38%
of endangered species), commercial developments (35%), water projects
(30%), outdoor recreation (27%), livestock grazing (22%), pollution (20%),
infrastructure and roads (17%), disruption of fire ecology (13%), and log-
ging (12%) (Stein et al. 2000; Wilcove and Master 2005).
As a result of farming, logging, and other human activities, very little
frontier forest—intact blocks of undisturbed forest large enough to support
all aspects of biodiversity—remains in many countries; the global decline
in frontier forest is estimated as approximately 0.5% per year during the
past decade (Hansen et al. 2013). In the Mediterranean region, which has
been densely populated by people for thousands of years, only 10% of the
original forest cover remains. An important point to remember here is that
wildlife individuals and populations are lost in approximate proportion
to the amount of habitat that has been lost; even though the Mediterra-
nean forest still exists in places, approximately 90% of the individuals and
populations of birds, butterflies, wildflowers, frogs, and mosses that once
existed are no longer there.
98 Chapter 4

(A) Biomes current and future


Temperate broadleaf and mixed forests

Tropical and subtropical dry broadleaf forests

Mediterranean forests, woodlands, and scrub

Temperate grasslands, savannas, and shrublands

Mangroves

Tropical and subtropical coniferous forests

Tropical and subtropical moist broadleaf forests

Temperate coniferous forests

Flooded grasslands and savannas

Tropical and subtropical grasslands, savannas, and shrublands

Montane grasslands and shrublands

Deserts and xeric shrublands


Converted
Boreal forests/taiga Development risk
Strictly protected
Tundra
0
0.0 0.2 0.4 0.6 0.8 1.0
Proportion of biome
(B) Ecoregions current development

Proportion converted (number of ecoregions)


>0.75 (117) 0.25–0.50 (143)
0.50–0.75 (118) <0.25 (359)

FIGURE 4.4 Many of the world’s major biomes have already had a large
proportion of their area converted to human uses. (A) Current and future conver-
sion. (B) Current conversion. (After Oakleaf et al. 2015.)
Threats to Biodiversity 99

Tropical rain forests


The destruction of tropical rain forests has come to be synonymous with the
rapid loss of species. The tropics occur between approximately 23.5° north
and south of the equator, and rain forests (sometimes written rainforests)
are those characterized by a closed, evergreen canopy often more than
25 m tall with abundant, woody vines and generally more than 2000 mm
rain per year (Turner 2001). Tropical rain forests occupy 7% of the Earth’s
land surface (Figure 4.5), but they are estimated to contain over 50% of
its species (Bradshaw et al. 2009; Corlett and Primack 2010; see Chapter 2).
Many tropical rain forest species are important to local economies and have
the potential for greater use by the entire world population (see Chapter
3). Tropical rain forests also have local significance as home to numerous
indigenous cultures, regional importance in protecting watersheds and
moderating climate, and global importance as sinks to absorb some of the
excess CO2 that is produced by the burning of fossil fuels. The original ex-
tent of tropical rain forests and related moist forests has been estimated at 17
million km2 based on current patterns of rainfall and temperature. A com-
bination of ground surveys, aerial photos, and remote-sensing data from
satellites has shown that only 11 million km2 remain (Corlett and Primack

Remaining tropical forest


Cleared tropical forest
Other areas designated as hotspots

FIGURE 4.5 The current extent of tropical forests, and the areas that have
been cleared of tropical forests. Note the extensive amount of land that has been
deforested in northern and southeastern South America, India, Southeast Asia,
Madagascar, and western Africa. The map also shows hotspots of biodiversity,
a subject that will be treated in further detail in Chapter 7. Many of the biodi-
versity hotspots in the temperate zone have a Mediterranean climate, such as
southwestern Australia, South Africa, California, Chile, and the Mediterranean
basin. This map is a Fuller Projection, which distorts the sizes and shapes of con-
tinents less than typical world maps. (Map created by Clinton Jenkins; originally
appeared in Pimm and Jenkins 2005.)
100 Chapter 4

2010). More than 60% of the recent loss has occurred in the Neotropics, with
Brazil alone accounting for almost half. Another third has occurred in Asia,
with Indonesia second to Brazil in the absolute rate of tropical forest loss.
Africa has contributed only 5.4% to the total area lost, reflecting the current
absence of industrial-scale agricultural clearance there. Strikingly, 55% of
all recent tropical forest losses occurred within only 6% of the total tropical
forest area, forming an “arc of deforestation” in the south and southeast
of the Brazilian Amazon, in much of Malaysia, and in Sumatra and parts
of Kalimantan in Indonesia. These areas are experiencing rapid deforesta-
tion; it is estimated that in Malaysia, 2% of the forest area is lost per year
(Hansen et al. 2013). Meanwhile, the rate of deforestation in the Brazilian
Amazon appears to be slowing down due to changes in government policy
and stricter enforcement of logging regulations.
In relative terms, the deforestation rate is greatest in Asia, averaging
about 1.2% per year, while in absolute terms, tropical America has the great-
est amount of deforestation because of its larger total area. If the current rates
of loss continue, there will be little tropical forest left after the year 2050,
except in the relatively small national parks and remote, rugged, or infertile
areas of the Amazon basin, Congo River basin, and New Guinea. The move
to establish large new parks in many tropical countries is cause for some
hope; however, these parks will need to be well funded and managed to
be effective in preserving biodiversity, as described in Chapter 8. In many
cases, these parks are only “paper parks” with few employees or facilities.
On a global scale, much of rain forest destruction may still result from
small-scale cultivation of crops by poor farmers, who are often forced onto
remote forest lands by poverty or sometimes moved there by government-
sponsored resettlement programs (Peres and Schneider 2012). Much of this
farming is shifting cultivation, a kind of subsistence farming sometimes
referred to as slash-and-burn, or swidden, agriculture, in which trees are
cut down and then burned. The cleared patches are farmed for two or three
seasons, after which soil fertility has usually diminished and soils have
eroded to the point where crop production is so low that the patches are
then abandoned and a new area is cleared (Phua et al. 2008). Although these
patches may recover with time, studies show that shifting cultivation has
a negative effect on both plants and animals (Mukul and Herbohn 2016).
Shifting cultivation is often practiced because farmers are unwilling or
unable to spend the time and money necessary to develop more permanent
forms of agriculture on land that they do not own and may not occupy for
very long. Rain forests are also destroyed by fuelwood production, mostly
to supply local villagers with wood for cooking fires. More than two bil-
lion people cook their food with firewood, so their impact is significant.
Increasing human populations in poor tropical countries will cause further
loss of tropical forests in coming decades.
In an increasing proportion of the tropics, however, clearance by peas-
ant farmers to meet subsistence needs is now dwarfed by clearance by large
landowners and commercial interests to create pasture for cattle ranching
or to plant cash crops, such as oil palms, soybeans, and rubber trees (Rosa
Threats to Biodiversity 101

FIGURE 4.6 Complex and diverse tropical forests give way to an African tea
plantation, a sea of green that supports almost no biodiversity—not only because
it is a monoculture, but also because of the liberal use of pesticides. Conversion
of tropical forests to agriculture is considered the most common cause of loss of
these ecosystems. This plantation and others like it were initially established to
be buffer zones around protected remnant forests; however, over time, the plan-
tations facilitated further encroachment and conversion by the growing local hu-
man population. In this region of western Kenya, almost 60% of the forest cover
was lost between 1984 and 2009, with 90% of this loss attributable to agriculture
(Cordeiro et al. 2015). (Photograph by Anna Sher.)

et al. 2012; Figure 4.6). Shifting cultivation is often less detrimental to


biodiversity than these large-scale, often pesticide-laden plantations in
which large areas are maintained under a uniform crop cover (Mukul
and Herbohn 2016). Commercial agriculture displaces poor farmers and
justifies the expansion of roads. In addition, large areas of rain forest are
damaged during commercial logging operations, most of which are poorly
managed selective logging. These logged forests are prone to widespread
fires due to the large numbers of branches and dead trees on the ground. In
many cases, logging operations precede conversion of tropical forest land
to agriculture and ranching. The relative importance of these enterprises
varies by geographical region: logging is a significant activity in tropical
Asia and America, cattle ranching is most prominent in tropical America,
and farming is more important for the rapidly expanding population in
tropical Africa (Corlett and Primack 2010). As tropical forests are affected
by people in any of these ways, their diversity of species often rapidly
declines (Gibson et al. 2011).
The destruction of tropical rain forests is frequently caused by demand
in industrialized countries for cheap agricultural products, such as rubber,
palm oil, cocoa, soybeans, orange juice, and beef, and for low-cost wood
102 Chapter 4

products (Nepstad et al. 2006). At present, most consumers in industrial-


ized countries are not aware of how their food choices affect land use.
Many people would be surprised to learn how widely palm oil is used in
processed food and consumer products, not to mention where it comes
from. It is also true that increasing proportions of these agricultural and
wood products are consumed within the countries that produce them or
exported to rapidly industrializing countries, such as China and India.
The story of Indonesian Borneo and Sumatra in Southeast Asia illus-
trates how rapid and serious rain forest destruction can be. Between 1990
and 2005, an incredible 42% of the lowland forest of these two large islands
was cleared. Most of the clearing was due to logging, both legal and il-
legal, and the development of cash crops, especially oil palms (Hansen et
al. 2009; Lee et al. 2014).

Other threatened habitats


The plight of the tropical rain forests is perhaps the most widely publicized
case of habitat destruction, but many other habitats are also in grave danger.

Tropical deciduous forests A month of less than 100 mm is con-


sidered dry in tropical forests, and trees that experience drought regularly
for long periods adapt by being deciduous (i.e., dropping their leaves). The
land occupied by tropical deciduous forests is more suitable for agricul-
ture and cattle ranching than is the land occupied by tropical rain forests.
These forests are also easier than rain forests to clear and burn. Moderate
rainfall in the range of 250–2000 mm per year allows mineral nutrients to
be retained in the soil, from which they can be taken up by plants. Conse-
quently, human population density is five times greater in deciduous for-
est areas of Central America than in adjacent rain forest areas. Today, the
Pacific coast of Central America has less than 2% of its original deciduous
forest remaining (WWF and McGinley 2009), and less than 3% remains in
Madagascar, which is home to the endangered lemurs, an endemic group
of primates (see Figure 5.10).

Grasslands Temperate grassland is another habitat type that has been


almost completely destroyed by human activities. It is relatively
easy to convert large areas of grassland to farmland or cattle
ranches. Between 1800 and 1950, as much as 98% of North Amer-
Between 1800 and 1950,
ica’s tallgrass prairie was converted to farmland. The remaining
as much as 98% of North
area of prairie is fragmented and widely scattered across the
America’s tallgrass
landscape. Increasing prices for agricultural products are cur-
prairie was converted to
rently driving the conversion of even marginal grasslands to
farmland. farmland (Wright and Wimberly 2013). Worldwide, only 4% of
temperate grasslands are protected (see Chapter 8).

Freshwater habitats Wetlands and other freshwater aquatic habitat


are critical habitats for fish, aquatic invertebrates, amphibians, and birds.
Threats to Biodiversity 103

They are also a resource for flood control, water filtration, and power pro-
duction (as described in Chapter 3). Freshwater systems are often filled in
or drained for development, or are otherwise altered by dams, channeliza-
tion of watercourses, and chemical pollution (Mitsch and Gosselink 2015).
Over half of the wetland ecosystems that existed in the early twentieth
century have been lost in North America, Europe, Australia, and China
(Moreno-Mateos et al. 2012). In the United States, 98% of the country’s 5.2
million km of streams have been degraded in some way to the point that
they are no longer considered wild or scenic. More importantly, their eco-
system functions and services are lost, including their ability to serve as
dispersal routes for aquatic animals and plants (see Chapter 3). Destruction
of wetlands and streams has been equally severe in other parts of the in-
dustrialized world, such as Europe and Japan. About 60%–70% of wetlands
in Europe have been lost. Only 2 of Japan’s 30,000 rivers can be considered
wild, without dams or some other major modification.
In the last few decades, major threats to wetlands and other aquatic
environments in developing countries have included massive development
projects involving drainage, irrigation, and dams, organized by govern-
ments and often financed by international aid agencies. The Three Gorges
Dam on the Yangtze River of China is a recent example (Sun et al. 2012;
Yang et al. 2014). The dam is the largest hydroelectric power plant in the
world, generating much-needed clean and renewable energy. However, the
dam and reservoir have displaced more than 1 million people, destroyed
untold numbers of ecosystems and archaeological sites, and altered the
river and delta systems, with unknown ecological consequences. The eco-
nomic benefits of such projects are important, but the rights of local people
and the value of ecosystems are often not adequately considered.

Marine coastal areas Human populations are increasingly concen-


trated in marine coastal areas. Twenty percent of marine coastal areas have
already been degraded, filled in, or highly modified by human activity,
despite their importance in the harvesting of fish, shellfish, seaweeds, and
other marine products (see Figure 3.5A). Coastal wetlands are also threat-
ened by pollution, dredging, sedimentation, destructive fishing practices,
invasive species, and now rising temperatures. Human impacts on these
habitats are less well studied than in the terrestrial environment, but they
are probably equally severe, especially in shallow coastal areas. Two coastal
habitats of special note are mangroves and coral reefs.

Mangroves Mangrove forests are among the most important wetland


communities in tropical areas (Polidoro et al. 2010), and it is estimated
that 2%–8% of global mangrove cover is lost each year (Miththapala 2008).
Composed of species that are among the few woody plants able to tolerate
salt water, mangrove forests occupy coastal areas with saline or brackish
water, typically where there are muddy bottoms. Such habitats, like salt
marshes in the temperate zone, are extremely important breeding grounds
104 Chapter 4

and feeding areas for shrimp and fish. In Australia, for example, two-
thirds of the species caught by commercial fishermen depend to some
degree on the mangrove ecosystem. Despite their great economic value
and their utility for protecting coastal areas from storms and tsunamis,
mangroves are often cleared for rice cultivation and commercial shrimp
and prawn hatcheries, particularly in Southeast Asia. Mangroves have also
been severely degraded by overcollection of wood for fuel, construction
poles, and timber throughout the region; it has been argued that there are
too few incentives for local users and managers to stop this overexploita-
tion (Máñez et al. 2014). Over half of the world’s mangrove ecosystems
have already been destroyed (Twilley and Day 2012). Today, almost 40%
of mangrove-dependent animal species are considered to be at high risk
of extinction (Daru et al. 2013).

Coral reefs Tropical coral reefs (see Chapter 2 opener) contain an es-
timated one-third of the ocean’s fish species in only 0.2% of its surface area
(Figure 4.7). At least 38% of all coral reefs have already been destroyed
(Butchart et al. 2010). A further 20% have been degraded by overfishing,
overharvesting, pollution, and the introduction of invasive species (MEA
2005). The most severe destruction is taking place in the Philippines, where
a staggering 90% of the reefs are dead or dying. In China, coral reefs have
declined by 80% over the past 30 years (Hughes et al. 2013). The main cul-
prits are pollution, which either kills the corals directly or allows excessive

Mediterranean Sea

Asia North Pacific North


Ocean America Atlantic
Red Sea
Ocean
Philippines Hawaii

Africa
Caribbean Sea
Equator
South
America
Indian
Ocean Australia
Great Barrier Reef
Madagascar
Critical; projected loss in 10–20 years South Pacific
Threatened; projected loss in 21–40 years Ocean

Stable

FIGURE 4.7 Extensive areas of coral are likely to be damaged or destroyed


by human activities over the next 40 years unless conservation measures can be
implemented. (After Bryant et al. 1998, with updates from Burke et al. 2011.)
Threats to Biodiversity 105

growth of algae; sedimentation resulting from deforestation, agriculture,


and coastal development; overharvesting of fish, clams, and other ani-
mals; and finally, stunning with dynamite or cyanide to collect the few
remaining living creatures for food and the aquarium trade. The effects
of global climate change, including ocean warming, ocean acidification,
and the increasing intensity of tropical storms fueled by warmer surface
waters, are also playing a major role in the rapid degradation of coral reefs
(Comeau et al. 2015; Schmidt et al. 2014). Worst of all is the combination of
local and global impacts, which makes it difficult or impossible for coral
reefs to bounce back from natural disturbances, such as hurricanes, to
which they may otherwise be relatively resilient.
Further losses of coral reefs are expected within the next 40 years, espe-
cially in tropical East Asia, the areas around Madagascar and East Africa,
and throughout the Caribbean. Elkhorn and staghorn corals, which were
formerly common in the Caribbean and gave three-dimensional structure
to the reef community, have already become rare in many locations (Car-
penter et al. 2009; Mora 2009).
Over the last 15 years, scientists have discovered extensive reefs of
corals living in cold water at depths of 300 m or more, many of which are
in the temperate zone of the North Atlantic. These coral reefs are rich in
species, many of which are new to science. Yet just as these communities
are being explored for the first time, they are being destroyed by trawlers,
which drag nets across the seafloor to catch fish. The trawlers destroy the
very coral reefs that protect and provide food for young fish. The damage
to these cold-water reefs by careless harvesting is costing the industry its
resource base in the long run.

Desertification
Many ecosystems in seasonally dry climates are degraded by human activi-
ties into human-made deserts, a process known as desertification (Lavauden
1927; Okin et al. 2009). These dryland ecosystems include grasslands, scrub,
and tropical deciduous forests as well as temperate shrublands, such as
those found in the Mediterranean region, southwestern Australia, South
Africa, central Chile, and California. Naturally occurring dry areas cover
about 41% of the world’s land area and are home to about one billion people.
Approximately 10%–20% of these drylands are at least moderately degraded,
and more than 25% of the productive capacity of their plant growth has
been lost (Sayre et al. 2013). These areas may initially support agriculture,
but their repeated cultivation, especially during dry and windy years, often
leads to soil erosion and loss of water-holding capacity in the soil. The land
may also be chronically overgrazed by domesticated livestock, and woody
plants may be cut down for fuel. Frequent fires during long dry periods
often damage the remaining vegetation. The result is the progressive and
largely irreversible degradation of the ecosystem and the loss of soil cover.
Ultimately, formerly productive farmland and pastures take on the appear-
106 Chapter 4

ance of a desert. Desertification has been ongoing for thousands of years in


the Mediterranean region and was known even to ancient Greek observers.
Worldwide, 9 million km 2 of arid lands have been converted to hu-
man-made deserts. These areas are not functional desert ecosystems, but
wastelands, lacking the flora and fauna characteristic of natural deserts.
The process of desertification is most severe in the Sahel region of Africa,
just south of the Sahara, where most of the native large mammal species
are threatened with extinction. The human dimension of the problem is
illustrated by the fact that the Sahel region is estimated to have 2.5 times
more people (100 million currently) than the land can sustainably support.
Further desertification appears to be almost inevitable, especially given
the higher temperatures and lower rainfall associated with predictions of
future climate change (Verstraete et al. 2009). In such areas, the solution
will be programs involving the implementation of sustainable agricultural
practices, the elimination of poverty, increased political stability, and con-
trol of the human population.

Habitat Fragmentation
In addition to being destroyed outright, habitats that formerly occupied wide,
unbroken areas are often divided into pieces by roads, fields, towns, and a
broad range of other human constructs. Habitat fragmentation is the process
whereby a large, continuous area of habitat is both reduced in area and di-
vided into two or more fragments (Figure 4.8). When habitat is destroyed,
a patchwork of habitat fragments may be left behind. These fragments are
often isolated from one another by a highly modified or degraded landscape,
and their edges experience altered conditions, referred to as edge effects,
such as increased wind, fire, species invasion, or predation (Murcia 1995;
Porensky and Young 2013). The fragments are often on the least desirable
land for human uses, such as steep slopes, poor soils, and inaccessible areas.
Fragmentation almost always occurs during a severe reduction in habi-
tat area, but it can also occur when habitat area is reduced to only a minor
degree by roads, railroads, canals, power lines, fences, oil pipelines, fire
lanes, or other barriers to the free movement of species. In many ways,
habitat fragments resemble islands of original habitat in an inhospitable,
human-dominated landscape. Habitat fragmentation is a serious threat
to biodiversity, as species are often unable to survive under the altered
conditions found in fragments.
Habitat fragments differ from the original habitat in three important
ways:
1. Fragments have a greater amount of edge per area of habitat (and thus
a greater exposure to edge effects).
2. The center of a habitat fragment is closer to an edge.
3. When a formerly continuous habitat hosting a large population is di-
vided into fragments, each fragment hosts a smaller population.
Threats to Biodiversity 107

14 ha

2 ha 2 ha

87 ha

FIGURE 4.8 The Biological Dynamics of Forest Fragmentation (BDFF) study


in Brazil is the largest and longest-running experiment of tropical forest frag-
mentation; forest fragments of different sizes were created 1980 when the land
was cleared for cattle ranches and farms. Shown in this aerial photograph are
forest fragments of 2 ha, 14 ha, and 87 ha. These fragments were surveyed for
species composition, microclimate, and other ecosystem characteristics before
the forest was cleared, and have been regularly surveyed over the past 36 years.
In addition, there have been forest blocks of equivalent size inside intact forest
that have been surveyed as controls. In the photograph, note that the fragments
vary in their distance from the intact forest. An important factor in the experi-
ment is that land use around the fragments has changed over time, with many of
the cattle pastures being abandoned, and undergoing succession to secondary
forests. For example, the left side of the 14 ha fragment is secondary vegetation
while the right side has recently been cleared. See Figure 4.10 for some of the
key results. (© DigitalGlobe.)

Threats posed by habitat fragmentation


Limits to dispersal and colonization Fragmentation may limit
a species’ potential for dispersal and colonization by creating barriers to
normal movements (Stouffer et al. 2011). In an undisturbed environment,
seeds, spores, and animals move passively or actively across the landscape.
Over time, populations of a species may build up and go extinct on a local
scale as the species disperses from one suitable site to another and the
biological community undergoes succession. At a landscape level, a series
of populations exhibiting this pattern of extinction and recolonization is
sometimes referred to as a metapopulation (see Figure 6.12).
When a habitat is fragmented, the potential for dispersal and colonization
is often reduced. Many bird, mammal, and insect species of the forest interior
108 Chapter 4

will not cross even short stretches of open ground (Laurance et al. 2009). If
they do venture into the open, they may find predators waiting on the forest
edge to catch and eat them. Agricultural fields 100 m wide may represent
an impassable barrier to the dispersal of many invertebrate species. Roads,
too, may be significant barriers to animal movement. Many species avoid
crossing roads, which represent an environment totally different from the
habitat they are leaving. For animals that do attempt to cross roads, motor
vehicles are a major source of mortality (Beebee 2013), as has been observed
in the endangered Florida panther (Kroll 2015). To deal with such problems,
highway officials are building animal underpasses, overpasses, and other
improvements to minimize animal mortality.
As species go extinct within individual fragments through
natural successional and metapopulation processes, new spe-
cies will be unable to arrive because of barriers to dispersal
The barriers that colonization, and the number of species present in the habitat
fragment a habitat fragment will decline over time. Extinction will be most rapid
reduce the ability of and severe in small habitat fragments.
animals to forage, find
mates, disperse, and Restricted access to food and mates Many animal
colonize new locations. species need to move freely across the landscape, either as indi-
Habitat fragmentation viduals or in social groups, to feed on widely scattered resources
often creates small (Becker et al. 2010). A given resource may be needed for only a
subpopulations that few weeks each year, or even only once in a few years, but when
are vulnerable to local a habitat is fragmented, species confined to a single habitat frag-
extinction. ment may be unable to migrate over their normal home range
in search of that scarce resource. Gibbons and other primates,
for example, typically remain in forests and forage widely for
fruits. Finding scattered trees with abundant fruit crops may be
crucial during episodes of fruit scarcity. Clearings and roads that break up
the forest canopy may prevent these primates from reaching nearby fruit-
ing trees because the primates are unable or unwilling to descend to the
ground and cross the intervening open landscape. Fences may prevent the
natural migration of large grazing animals such as wildebeest and bison,
forcing them to overgraze unsuitable habitat, which eventually leads to
starvation and further degradation of the habitat (Gates et al. 2012).
Barriers to dispersal can also restrict the ability of widely scattered
species to find mates, leading to a loss of reproductive potential for many
animal species. Plants may have reduced seed production if butterflies
and bees are less able to migrate among habitat fragments to pollinate
flowers.

Creation of smaller populations Habitat fragmentation may


precipitate population decline and extinction by dividing an existing wide-
spread population into two or more subpopulations in restricted areas.
These smaller populations are then more vulnerable to inbreeding de-
Threats to Biodiversity 109

pression, genetic drift, and other problems associated with small popu-
lation size (see Chapter 5). While a large area may support a single large
population, it is possible that none of the smaller subpopulations in the
fragments will be sufficiently large to persist, even if the total area is the
same. Connecting the fragments with properly designed movement cor-
ridors may be the key to maintaining populations.

Interspecies interactions Habitat fragmentation increases the


vulnerability of the fragments to invasion by nonnative and native pest
species (Aguirre-Acosta et al. 2014; Flory and Clay 2009). Road edges them-
selves may represent dispersal routes for invasive species. The forest edge
represents a high-energy, high-nutrient, disturbed environment in which
many pest species of plants and animals can increase in number and then
disperse into the interior of the fragment.
Omnivorous animals may increase in population size along forest
edges, where they can eat foods, including eggs and nestlings of birds,
from both undisturbed and disturbed habitats. In the coniferous forests
of Finland, predation on nests was found to be higher in edges of clear-cut
areas than in either the forest interior or forest corridors, and sometimes
higher than in the clear-cut area (Huhta and Jokimäki 2015). The invading
predators include not only native omnivores—such as raccoons, skunks,
and blue jays in North America—but also introduced domesticated cats.
In a study done in the southeastern United States, successful hunting cats
were found to capture an average of 2.4 prey items during 7 days of roam-
ing, with Carolina anoles (Anolis carolinensis) being the most common prey
species (Loyd et al. 2013). In rural Poland, an examination of both scat and
stomach contents and prey brought home by cats found mammals to be
the most common prey, followed by birds (Krauze et al. 2012).
Nest-parasitizing cowbirds, which live in fields and edge habitats, use
habitat edges as invasion points, flying up to 15 km into forest interiors,
where they lay their eggs in the nests of forest songbirds (Cox et al. 2012;
Lloyd et al. 2005). The combination of habitat fragmentation, increased
nest predation, and destruction of tropical wintering habitats is probably
responsible for the dramatic decline of certain migratory songbird spe-
cies of North America, such as the cerulean warbler (Dendroica cerulea),
particularly in the eastern half of the United States (Valiela and Martinetto
2007; With 2015). Populations of deer and other herbivores can also build
up in edge areas, where plant growth is lush, eventually overgrazing the
vegetation and selectively eliminating certain rare and endangered plant
species for several kilometers into the forest interior.
Habitat fragmentation also puts wild populations of animals in closer
proximity to domesticated animals. Diseases of domesticated animals can
then spread more readily to wild species, which often have no immunity
to them. There is also the potential for diseases to spread from wild spe-
cies to domesticated plants, animals, and even people, once the level of
110 Chapter 4

contact increases. Fragmented forest habitats characteristic of suburban


development often have high densities of white-footed mice and black-
legged ticks with high rates of infection with Lyme disease, along with a
corresponding increase in Lyme disease in people living in those areas
(Tran and Waller 2013).

Edge effects
Habitat fragmentation greatly increases the amount of edge relative to the
amount of interior habitat. A simple example will illustrate these characteris-
tics and the problems they can cause. Consider a square conservation reserve
1000 m (1 km) on each side (Figure 4.9). The total area of the reserve is 1 km2
(100 ha). The perimeter (or edge) of the reserve totals 4000 m. A point in the
middle of the reserve is 500 m from the nearest perimeter. If the principal
edge effect for birds in the reserve is predation by domesticated cats and
introduced rats, which forage 100 m into the forest from the perimeter of the
reserve and prevent forest birds from successfully raising their young, then
only the reserve’s interior—64 ha—is available to the birds for breeding. If
the reserve is divided into four fragments by a road and a railroad, each of
which is 495 m in area, the nesting habitat is further reduced to 8.7 ha in each
fragment, for a total of 34.8 ha. Even though the road and railroad remove
only 2% of the reserve area, they reduce the habitat available to the birds
by about half. The implications of edge effects can be seen in the decreased

(A) (B)

8.7 ha 8.7 ha

1000 m Interior = 64 ha Railroad


(1 km)

8.7 ha 8.7 ha

100 800 m 100 Road


m m
1000 m Interior = 8.7 ha × 4 = 34.8 ha

FIGURE 4.9 A hypothetical example shows how habitat area is reduced by


fragmentation and edge effects. (A) A 1 km2 protected area. Assuming that edge
effects (gray) penetrate 100 m into the reserve, approximately 64 ha are available
as usable habitat for nesting birds. (B) The bisection of the reserve by a road and
a railroad, although taking up little actual area, extends the edge effects so that
almost half the breeding habitat is destroyed. Edge effects are proportionately
greater when forest fragments are irregular in shape, as is usually the case.
Threats to Biodiversity 111

ability of birds to live and breed in small forest fragments com-


pared with larger blocks of forest habitat. Habitat fragmentation
The microenvironment at a forest fragment edge is different increases edge effects—
from that in the forest interior. Some of the more important dif- changes in light,
ferences are greater fluctuations in levels of light, temperature, humidity, temperature,
humidity, and wind (Laurance et al. 2011). These edge effects and wind that may be
are most evident up to 100 m inside the forest, although certain less favorable for many
effects are detectable up to 400 m from the forest edge (Figure species living there.
4.10). Because so many plant and animal species are precisely
adapted to certain levels of temperature, humidity, and light,
changes in those conditions eliminate many species from for-
est fragments. Shade-tolerant wildflower species of temperate
forests, late-successional tree species of tropical forests, and

Increased wind disturbance

Elevated tree mortality

Altered species composition of leaf-litter invertebrates

Altered abundance and diversity of leaf-litter invertebrates

Altered height of greatest foliage density

Reduced canopy height

Faster recruitment of disturbance-adapted trees


Edge effect

Reduced soil moisture

Lower canopy-foliage density Emergent

Increased air temperature

Reduced understory-bird abundance Canopy

Lower relative humidity Understory

Higher understory-foliage density Shrub


Lower soil-moisture content Floor (soil, leaf litter)

Reduced density of fungal fruiting bodies

100 200 300 400 500


Edge penetration distance (m)

FIGURE 4.10 Edge effects in the Amazon rain forest as measured in the ex-
periment shown in Figure 4.8. The bars indicate how far into the forest fragment
the specified effect occurs. For example, trees growing within 300 m of an edge
have a higher mortality rate, and the average height of trees in the forest canopy
(see drawing) is reduced within 100 m of the edge. (After Laurance et al. 2002.)
112 Chapter 4

humidity-sensitive animals such as amphibians are often rapidly elimi-


nated by habitat fragmentation.
When a forest is fragmented, increased wind, lower humidity, and
higher temperatures make fires more likely (Uriarte et al. 2012). Fires may
spread into habitat fragments from nearby agricultural fields that are being
burned regularly, as in sugarcane harvesting, or from the irregular activi-
ties of farmers practicing slash-and-burn agriculture. Forest fragments may
be particularly susceptible to fire damage when wood has accumulated
at the edge of the forest where trees have died or have been blown down
by the wind. In Borneo and the Brazilian Amazon, millions of hectares of
tropical moist forest burned during unusually dry periods in 1997 and 1998.
A combination of factors contributed to these environmental disasters:
forest fragmentation due to road construction, farming, the accumulation
of brush following selective logging, and human-caused fires.

Environmental Degradation and Pollution


Even when a habitat is unaffected by overt destruction or fragmentation,
the ecosystems and species in that habitat can be profoundly affected by
human activities. Ecosystems can be damaged and species driven to ex-
tinction by external factors that do not obviously change the structure of
the dominant plants or other features in the community. Sometimes this
damage and its cause(s) are quite obvious (Figure 4.11). For example,
keeping too many cattle in a grassland community gradually damages
it, often eliminating many native species and favoring invasive species
that can tolerate grazing and trampling. On the other hand, out of sight
from the public, fishing trawlers drag across an estimated 15 million km2
of ocean floor each year, an area 150 times greater than the area of forest
cleared in the same period. The trawling destroys delicate creatures such
as anemones and sponges, reduces species diversity and biomass, and
alters community structure (Hinz et al. 2009; Maynou and Cartes 2012).
Other types of environmental damage are not visually apparent even
though they occur all around us, every day, in nearly every part of the
world. The most subtle and universal form of environmental degradation
is pollution, commonly caused by pesticides, herbicides, sewage, fertilizers
from agricultural fields, industrial chemicals and wastes, emis-
sions from factories and automobiles, and sediment deposits
from eroded hillsides. The general effects of pollution on water
Pollution of the air, quality, air quality, and even the global climate are cause for
water, and soil by great concern, not only because of their threats to biodiversity,
chemicals, wastes, and but also because of their effects on human health and agricul-
the by-products of energy ture (Dearborn and Kark 2010). Although environmental pol-
production destroys lution is sometimes highly visible and dramatic, as in the case
habitats in insidious of massive oil spills, the subtle, unseen forms of pollution are
ways. probably the most threatening—primarily because they are so
insidious.
Threats to Biodiversity 113

FIGURE 4.11 This 80-mile-long spill of toxic materials from the Gold King
Mine, which extended through three states and two Native American reserva-
tions in 2015, prompted a state of emergency in the southwestern United States.
The extent of the harm done to wildlife and people by the arsenic, lead, mercury,
and cadmium released is not fully known. An EPA report said that human error,
in the form of a botched cleanup effort, was to blame for the 3-million-gallon
spill (Denver Post 2016). (Animas River at Bakers bridge near Durango, Colorado,
USA, taken August 6, 2015. © Whit Richardson/Alamy Stock Photo.)

Pesticide pollution
The dangers of pesticides were brought to the world’s attention in 1962
by Rachel Carson’s influential book Silent Spring (see Figure 1.4). Carson
described a process known as biomagnification, through which dichloro-
diphenyl-trichloro-ethane (DDT) and other organochlorine pesticides be-
come concentrated as they ascend the food chain (Kohler and Triebskorn
2013; Weis and Cleveland 2008; Figure 4.12). These pesticides, used on
crop plants to kill insects and sprayed on water bodies to kill mosquito
larvae, were harming wildlife populations, especially birds, such as hawks
and eagles, that eat large amounts of insects, fish, or other animals exposed
to DDT and its by-products.
Recognition of this situation in the 1970s led many industrialized coun-
tries to ban the use of DDT and other chemically related pesticides. The ban
eventually allowed the partial recovery of many bird populations (Figure
4.13). Nevertheless, massive use of pesticides—even DDT—persists because
of their benefits to people. For example, DDT is still highly effective in
controlling mosquitoes, which, through the malaria they spread, are still
a significant cause of death in tropical regions. These benefits must be
114 Chapter 4

Humans

Marine mammals Bald eagles

Large fish that


Sharks eat small fish

Seabirds

Pacific salmon

Fish that eat zooplankton Tufted puffin

Pacific herring and sand lance

Juvenile fishes and other zooplankton Algae

DDT, pesticides,
PCBs, mercury,
and other toxic
chemicals

FIGURE 4.12 Toxic chemicals become successively concentrated at higher


levels in the food chain, leading to health problems for humans, marine mam-
mals, seabirds, and raptors. (After Groom et al. 2006.)

weighed against harm not only to endangered animal species, but also to
people, particularly the workers who handle these chemicals in the field
and the consumers of the agricultural products, such as crops and even
chicken eggs, exposed to these chemicals (Bouwman et al. 2015). Even
plants, animals, and people living far from where the chemicals are ap-
plied can be harmed; high concentrations of pesticides are found even in
the tissues of polar bears in northern Norway and Russia, where they have
Introduction to Conservation Biology 1E Primack/Sher
a harmful effect on bear health (Elliott and Elliott 2013).
Sinauer Associates
Morales Studio/SA
Primack_Sher1E_04.12 Water
Date pollution
12-17-15
Water pollution has negative consequences for all species: it destroys im-
portant food sources and contaminates drinking water with chemicals
that can cause immediate and long-term harm to the health of humans and
other species that come into contact with the polluted water (Feist et al.
2011) (see Figure 4.11). In the broader picture, water pollution often severely
Threats to Biodiversity 115

Breeding range

4
peregrine falcon pairs (thousands)
Estimated number of nesting

FIGURE 4.13 Peregrine falcons are


2 now breeding in many areas across North
America. Their population declined dra-
matically when DDT use began in the
1940s and then recovered after DDT was
1
banned in 1972 (Hoffman and Smith 2003).
(After Canadian Wildlife Service and Con-
0 necticut Department of Environmental
1900 1920 1940 1960 1980 2000 Protection; photograph courtesy of the US
Year Fish and Wildlife Service.)

damages aquatic ecosystems. Rivers, lakes, and oceans are sometimes used
as open sewers for industrial wastes and residential sewage. And higher
densities of people almost always mean greater levels of water pollution.
Pesticides, herbicides, petroleum products, heavy metals (such as mer-
cury, lead, and zinc), detergents, toxic chemicals such as polychlorinated
biphenyls (PCBs), and industrial wastes directly kill organisms such as in-
sect larvae, fish, amphibians, and even marine mammals living in aquatic
environments. Pollution is a threat to 90% of the endangered fishes and
freshwater mussels in the United States. An increasing source of pollution
in coastal areas is the discharge of nutrients and chemicals from shrimp
and salmon farms. Medicines used by people or given to domesticated
animals can enter the aquatic environment through sewage, either because
waste treatment
Introduction plants
to Conservation cannot
Biology not remove them or because they leak into
1E Primack/Sher
wells
Sinauer(Schaider
Associates et al. 2014). These biologically active chemicals, especially
Morales Studio/SA
hormones, can have an adverse effect on the physiology, behavior, and re-
Primack_Sher1E_04.13 Date 12-17-15 1/15/16
production of fish and other animals that ingest them (Brodin et al. 2013).
In contrast to wastes in the terrestrial environment, which have primar-
ily local effects, toxic wastes in aquatic environments diffuse over a wide
area. Toxic chemicals, even at very low concentrations in the water, can
be lethal to aquatic organisms through biomagnification. Many aquatic
environments are naturally low in essential minerals, such as nitrates and
phosphates, and aquatic plant and animal species have adapted to their
natural absence by developing the ability to process large volumes of water
116 Chapter 4

and concentrate these minerals. When these species process polluted water,
they concentrate toxic chemicals along with the essential minerals, and
the toxins may eventually poison them. Species that feed on these aquatic
species ingest the toxic chemicals they have concentrated. One of the most
serious consequences for humans is the accumulation of mercury and
other toxins by long-lived predatory fishes, such as swordfish and sharks,
and its effect on the nervous systems of people who eat these types of fish
frequently (Jaeger et al. 2009).
Even essential minerals that are beneficial to plant and animal life can
become harmful pollutants at high concentrations (McWilliams 2013). An-
thropogenic releases of human sewage, agricultural fertilizers, detergents,
and industrial wastes often add large amounts of nitrates and phosphates
to aquatic systems, initiating the process of eutrophication. Humans re-
lease as much nitrate into the environment as is produced by all natural
processes, and the anthropogenic release of nitrogen is expected to keep
increasing as the human population continues to increase. Even small
amounts of these nutrients can stimulate growth, and high concentrations
often result in thick “blooms” of algae at the surfaces of ponds and lakes
(Figure 4.14). These algal blooms may be so dense that they outcompete
other plankton species and shade bottom-dwelling plant species. As the
algal mat becomes thicker, its lower layers sink
to the bottom and die. The bacteria and fungi
that decompose the dying algae multiply in
response to this added sustenance and con-
sequently absorb all the oxygen in the water.
Without oxygen, much of the remaining animal
life dies off, sometimes visibly in the form of
masses of dead fish floating on the water’s sur-
face. The result is a greatly impoverished and
simplified community, a dead zone consisting of
only those species that are tolerant of polluted
water and low oxygen levels.
This process of eutrophication can also affect
marine systems, particularly coastal areas and
bodies of water in confined areas, such as the
Gulf of Mexico, the Mediterranean, the North
and Baltic Seas in Europe, and the enclosed seas
of Japan which have large anthropogenic inputs
of nutrients (Greene et al. 2009). In warm tropi-
cal waters, eutrophication favors algae, which
grow over coral reefs and completely change

FIGURE 4.14 Large algal blooms such as this


one on a pond in Oulu, Finland, are often the result
of human activities, and when extensive can have
significant negative effects on an entire ecosystem.
(© RJH_CATALOG/Alamy Stock Photo.)
Threats to Biodiversity 117

the reef community (Díaz-Ortega and Hernández-Delgado 2014). The key


to stopping eutrophication and its negative effects is to reduce the release
of excess nutrients through improved sewage treatment and better farming
practices, including reduced applications of fertilizer and establishment
of buffer zones between fields and waterways.
Sediments eroded from logged or farmed hillsides can also harm aquat-
ic ecosystems. The sediment covers submerged plant leaves and other green
surfaces with a muddy film that reduces light availability and can block O2
and CO2 exchanges, thus decreasing the rate of photosynthesis. Increasing
water turbidity (a decrease in water clarity) reduces the depth at which
photosynthesis can occur and may prevent animal species from seeing,
feeding, and living in the water. Sediment loads are particularly harmful
to the many coral species that require crystal-clear waters to survive.

Air pollution
In the past, people assumed that the atmosphere was so vast that materi-
als they released into the air would be widely dispersed and their effects
would be minimal. But today, several types of air pollution are so wide-
spread that they have damaged whole ecosystems. These same pollutants
also have severe effects on human health, demonstrating again the com-
mon interests shared by people and nature.

Acid rain Acid rain is produced when industries such as Acid rain and other
smelting operations and coal- and oil-fired power plants release effects of air pollution
huge quantities of nitrogen oxides and sulfur oxides into the air, are increasing rapidly in
where those chemicals combine with moisture in the atmosphere Asian countries as they
to produce nitric and sulfuric acids. These acids are incorporated industrialize. Acid rain
into cloud systems and dramatically lower the pH (the standard is particularly harmful to
measure of acidity) of rainwater, leading to the weakening and freshwater species.
death of trees over wide areas. Acid rain also lowers the pH of
soil moisture and water bodies and increases the concentration
of toxic metals such as aluminum in the soil and water.
Increased acidity alone damages many plant and animal species; as the
acidity of water bodies increases, many fish either fail to spawn or die out-
right. Increased acidity, along with both aquatic and terrestrial pollution, is
a contributing factor to the dramatic decline of many amphibian popula-
tions throughout the world (Brühl 2013; Hayes et al. 2010). Most amphibian
species depend on bodies of water for at least part of their life cycle, and
a decline in water pH causes a corresponding increase in the mortality
of eggs and young animals. Acidity also inhibits the microbial process
of decomposition, lowering the rate of mineral recycling and ecosystem
productivity. Many ponds and lakes in industrialized countries have lost
large portions of their animal communities as a result of acid rain. These
damaged water bodies are often in supposedly pristine areas hundreds of
kilometers from major sources of urban and industrial pollution, such as
the North American Rocky Mountains and Scandinavia. While the acidity
of rain is decreasing in many areas because of better pollution control, acid
118 Chapter 4

rain remains a serious problem. In developing countries such as China,


acid rain is becoming a greater problem as the country powers its rapid
industrial development with fuels that are high in sulfur.

Ozone production and nitrogen deposition Automobiles,


power plants, and industrial activities release hydrocarbons and nitrogen ox-
ides as waste products. In the presence of sunlight, these chemicals react with
the atmosphere to produce ozone and other secondary chemicals, collectively
called photochemical smog. Although ozone in the upper atmosphere is im-
portant in filtering out ultraviolet radiation, high concentrations of ozone at
ground level damage plant tissues and make them brittle, harming biological
communities and reducing agricultural productivity. Ozone and smog are
detrimental to people and other animals when inhaled, so both people and
biological communities benefit from air-pollution controls. Smog can be so
severe that people may avoid outdoor activities. Severe smog is becoming
an increasing problem in China and other Asian countries.
When airborne nitrogen compounds are deposited in terrestrial envi-
ronments by rain and dust, ecosystems throughout the world are damaged
and altered by potentially toxic concentrations of nitrogen, and many spe-
cies are unable to survive in the altered conditions (Bobbink et al. 2010).
In particular, the combination of nitrogen deposition and acid rain is re-
sponsible for a decline in the density of soil fungi that form beneficial
relationships with trees (see Chapter 3).

Toxic metals Leaded gasoline (still used in many developing countries,


despite its clear danger to human health), coal burned for heat and power,
mining and smelting operations, and other industrial activities release large
quantities of lead, zinc, mercury, and other toxic metals into the atmosphere
(Wade 2013). These elements are directly poisonous to plant and animal life
and can cause permanent injury to children (see Figure 4.11). The effects of
these toxic metals are particularly evident in areas surrounding large smelt-
ing operations, where there may be measurable negative effects on life for
miles around.
Enforcement of local and national policies and regulations may some-
times reduce levels of air pollution; eliminating lead from gasoline is one
such example. Concentrations of air pollutants are declining in certain
areas of North America and Europe, but they continue to rise in many other
areas of the world that are undergoing industrialization. Because air and
water pollutants do not observe political boundaries, international agree-
ments addressing these issues are important, as we will see in Chapter 11.
Lowering levels of air and water pollution is necessary for the health of
both human populations and biodiversity.

Global Climate Change


Carbon dioxide, methane, and other trace gases in the atmosphere allow
light energy to pass through and warm the surface of the Earth. However
Threats to Biodiversity 119

these gases, as well as water vapor (in the form of clouds), are able to trap
the energy radiating from the Earth as heat, slowing the rate at which
heat leaves the Earth’s surface and radiates back into space. These gases
are called greenhouse gases because they function much like the glass
in a greenhouse, which is transparent to sunlight but traps energy inside
the greenhouse once it is transformed into heat. The similar warming ef-
fect of Earth’s atmospheric gases is called the greenhouse effect. We can
imagine these gases as “blankets” over the Earth’s surface: the denser the
concentration of gases, the more heat is trapped near the Earth, and the
higher the planet’s surface temperature.
The greenhouse effect allows life to flourish on Earth—without it, the
temperature at the Earth’s surface would fall dramatically. Today, however,
as a result of human activities, concentrations of greenhouse gases are
increasing so much that they are already affecting the Earth’s climate (Gore
2006; IPCC 2014). The term global warming is used to describe the rise in
temperatures resulting from the increase in greenhouse gases, and global
climate change refers to the complete set of climate characteristics that are
changing now and will continue to change in the future because of this
increase, including patterns of precipitation and wind.
During the past 130 years, global atmospheric concentrations of carbon
dioxide, methane, and other trace gases have been steadily increasing,
primarily as a result of the burning of fossil fuels—coal, oil, and natural
gas (Climate Central 2012; IPPC 2013). Clearing of forests by logging and
for agriculture also contribute to rising concentrations of CO2. Through
all these activities, humans currently release about 70 million tons of CO2
into the atmosphere every day. The concentration of CO2 in the atmosphere
has increased from 290 parts per million (ppm) to around 400 ppm over
the last 100 years, and it is projected to reach 580 ppm at some point in the
latter half of the twenty-first century.
There is broad scientific agreement among the Intergovernmental Panel
on Climate Change (IPCC), a study group of leading scientists
organized by the United Nations, that the increased levels of
greenhouse gases have affected the world’s climate and ecosys-
There is a broad consen-
tems already and that these effects will increase in the future
sus among scientists that
(IPCC 2013). An extensive review of the evidence supports the
conclusion that global surface temperatures have increased by increased atmospheric
0.8°C (1.4°F) during the last 110 years (Figure 4.15). In fact, Au- concentrations of car-
gust 2015 was the hottest month in Earth’s recorded history; the bon dioxide and other
average global temperature was 15.6°C (60.1°F), 0.88°C (1.58°F) greenhouse gases
above the twentieth-century average, while temperatures in the produced by human
Pacific Ocean were 2°C (3.6°F) above the 1981–2010 average in activities have already
the eastern half of the equatorial Pacific Ocean (NOAA 2015). resulted in warmer
The El Niño (a periodic warming of Pacific temperatures) dur- temperatures and will
ing winter of 2015–2016 is one of the most severe in at least 50 continue to affect Earth’s
years, creating record highs in several cities during December climate in the coming
(NOAA 2016). Temperatures at high latitudes, such as in Siberia, decades.
Alaska, and Canada, have increased more than in other regions.
120 Chapter 4

Greatest temperature
increases in northern
polar latitudes

Least temperature
increases over
open ocean
Change from average mean surface temperature 1980–1999 (ºC)
+1.0 +2.0 +3.0 +4.0 +5.0 +6.0
+1.5 +2.5 +3.5 +4.5 +5.5

FIGURE 4.15 Over the last 130 years, atmospheric CO2 concentrations have in-
creased dramatically as a result of human activities, and global temperatures have
risen as a result. Here, the average annual temperature used for comparison is that
for the period from 1980 to 1999; temperature changes are reported in terms of dif-
ference (anomaly) from this average annual temperature. Global annual tempera-
tures were colder than average prior to 1980, when annual temperatures began to
be warmer than average. (After Karl 2006, updated from NOAA 2015.)

Climatologists predict that the world climate will warm by an addi-


tional 1°C–3°C (1.8°F–5.4°F) by 2100 as a result of increased atmospheric
concentrations of CO2 and other gases. The increase could be even greater,
in the range of 3°C–6°C (5.4°F–10.8°F), if CO2 levels rise faster than pre-
dicted; conversely, it could be slightly less if all countries reduce their
emissions of greenhouse gases in the very near future. The increase in
temperature will be greatest at high latitudes and over large continents
(IPCC 2013; Figure 4.16). Rainfall has already started to increase on a

FIGURE 4.16 (A) All climate models predict that the greatest warming will
take place in the northern polar regions. The graphs show satellite data-derived
Introduction to Conservation Biology 1E Primack/Sher
measurements of the extent of polar sea ice from 1979 to 2013 for the Arctic
Sinauer Associates
(top), Antarctica (middle), and the global combined total (bottom). Straight lines
Morales Studio/SA
Primack_Sher1E_04.15 represent the overall trend of observed data. Slight gains in polar ice in the south
Date 12-17-15
do not compensate for dramatic losses in the north. (B) Polar ice caps are already
melting at alarming rates, as these walrus crowded onto an ice floe in the Bering
Sea off Alaska seem to attest. (A, Graphs by Josh Stevens, NASA Earth Observa-
tory; B, photograph by Budd Christman, courtesy of NOAA.)
Threats to Biodiversity 121

(A)
Arctic
1.0

0.0

–1.0
Monthly deviations in sea ice extent (1,000,000 km2)

–2.0

Antarctic
1.0

0.0

–1.0

–2.0

Combined
1.0

0.0

–1.0

–2.0

1980 1985 1990 1995 2000 2005 2010


Years

(B)
122 Chapter 4

global scale and will continue to do so, but changes in rainfall will vary
by region, with some regions showing decreases. There will also probably
be an increase in extreme weather events, such as hurricanes, floods,
snowstorms, and regional drought, associated with global warming. In
tropical deciduous forests and savannas, warmer conditions will result in
an increased incidence of fire. In coastal areas, storms will cause increased
destruction of cities and other human settlements and will severely dam-
age coastal vegetation, including beaches and coral reefs. The series of
hurricanes that devastated the southern United States in 2005, including
Hurricane Katrina, could be an indication of what the future may bring.
As a result of global climate change, climate regions in the
northern and southern temperate zones will be shifted toward
the poles. This warming is expected to create a snowball effect
As rainfall patterns in which thawing tundra and melting permafrost, the layer of
change and most regions soil that typically stays frozen throughout the year, will increase
become warmer, many the rate of CO2 release through respiration by soil microbes,
terrestrial plant and but more research is needed to understand its extent (Schuur
animal species may not et al. 2015). However, the effects of warming are already being
be able to adapt quickly observed both at high altitudes and at high latitudes. Alpine
enough to survive. plants are found growing higher on mountains and blooming
Climate change may also earlier in the spring; Rocky Mountain species are blooming as
have huge impacts on much as a month earlier now than a century ago (Munson and
marine ecosystems and Sher 2015), which puts them at potential risk of frost damage and
coastal areas occupied by disconnects them from their pollinators (Inouye et al. 2002). Mi-
people. grating birds have been observed spending longer times at their
summer breeding grounds (Bussière et al. 2015). In the coming
century, global climate change is predicted to have profound
effects on Arctic, boreal, and alpine ecosystems as a result of
warmer temperatures and a longer growing season.
The effects of global climate change on temperature and rainfall are
also expected to have dramatic effects on tropical ecosystems (IPCC 2014).
Many tropical species and biological communities appear to have nar-
row tolerances for temperature and rainfall, so even small changes in the
climate could have major effects on species composition, cycles of plant
reproduction, patterns of migration, and susceptibility to fire (Corlett 2011).
Major contractions in the area of rain forest are quite likely. In particular,
cool-adapted species that live atop tropical mountains could be highly
vulnerable to increasing temperatures; as bands of vegetation move higher
on mountains, the species at the top will have nowhere to go (Şekercioğlu
et al. 2012).

Ocean acidification, warming, and rising sea level


Evidence indicates that ocean water temperatures have increased over the
last 40 years by about 0.44°C. As a consequence, certain marine species are
expanding their ranges farther from the equator, and coral reefs and other
marine habitats are threatened by rising seawater temperatures (Pandolfi
et al. 2011). In the coastal waters off California, warm-water southern spe-
Threats to Biodiversity 123

cies are increasing in abundance while cold-water northern species are


declining. Zooplankton are declining in some areas because of warmer
seawater temperatures, with dire consequences for the marine animals that
feed on them (Robinson et al. 2008). Abnormally high water temperatures
in oceans across the globe in 1998 caused corals to sicken and expel the
symbiotic algae that would normally live inside their bodies and provide
them with essential carbohydrates. These “bleached” corals then suffered
a massive dieback, with an estimated 70% coral mortality in Indian Ocean
reefs in 1998. Although reefs in some places have been able to recover from
that event (Gilmour et al. 2013), subsequent global bleaching events in 2010
and 2015 have taken a significant toll (Figure 4.17).
Increased anthropogenic CO2 in the atmosphere has also caused an-
other problem: as much as a third of it is absorbed by our oceans, causing
a reduction in the pH (Doney et al. 2009). Ocean acidification is associated
with many problems, most notably the decrease in calcium carbonate satu-
ration states that affect shell-forming marine animals.
Warming temperatures are already causing mountain glaciers to melt
and the polar ice caps to shrink, and this process will continue and ac-

FIGURE 4.17 The 2015 coral bleaching event had dramatic effects on this reef in
American Samoa. The photo on the left was taken in December 2014, and the one on
the right was taken in February 2015. The 2015 event was the third global bleaching
event in recorded history, the others taking place in 2010 and 1998; all three events
coincided with periods of abnormally high ocean temperatures. The photos were
taken by the XL Catlin Seaview Survey, a group documenting bleaching across the
globe. Although fluctuations in ocean temperature are normal, the geographic pat-
tern and intensity of these recent warming events have been very unusual.
124 Chapter 4

celerate. As a result of this melting and the thermal expansion of ocean


water, sea level is currently rising at a rate of about 3 cm (just over an inch)
per decade. By the year 2100, sea level is predicted by the IPCC to rise by
40–60 cm (16–24 in.), and possibly by as much as 100 cm (3 ft) (IPCC 2013;
Jones 2013). Many scientists, however, consider these predictions to be too
conservative and believe that a rise of up to 130 cm (4 ft) is possible. In
regions with low elevational gradients, a 4-ft rise in sea level could affect
areas thousands of meters inland; massive flooding of coastal cities and
communities will certainly occur. Much of the current land area of low-
lying countries such as Bangladesh could be underwater within 100 years.
Somewhere between 25% and 80% of coastal wetlands could be altered by
rising sea level. Many coastal cities, such as Miami and New York, will
have to build expensive seawalls or be flooded by the rising waters. There
is evidence that this process has already begun; sea level has already risen
by about 20 cm (8 in.) over the last 100 years (IPCC 2013).

The overall effect of global warming


Global climate change has the potential to radically restructure ecosystems
by changing the ranges of many species. The pace of this change could
overwhelm the natural dispersal abilities of species. There is mounting
evidence that this process has already begun (Table 4.2). The distributions
of bird and plant species are moving poleward, and reproduction is oc-

TABLE 4.2 Some Effects of Global Warming


1. Increased temperatures and incidence of heat waves
Examples: 2015 was the hottest year worldwide in the historical record so
far; previously, the warmest year was 2014. An August 2003 heat wave in
France killed over 10,000 people as temperatures reached 40°C (104°F).
2. Melting of glaciers and polar ice
Examples: Arctic Ocean summer ice has declined in area by 15% over the
past 25 years. Since 1850, glaciers in the European Alps have disappeared
from more than 30%–40% of their former range.
3. Rising sea level
Example: Since 1938, one-third of the coastal marshes in a wildlife refuge in
the Chesapeake Bay have been submerged by rising seawater.
4. Earlier spring activity
Example: One-third of English birds are now laying eggs earlier in the year,
and two-thirds of temperate plant species are now flowering earlier than
they did several decades ago.
5. Shifts in species ranges
Example: Two-thirds of European butterfly species studied are now found 35
to 250 km farther north than recorded several decades ago.
6. Population declines
Example: Adélie penguin populations have declined over the past 25 years as
their Antarctic sea ice habitat melts away.
Sources: Data from Union of Concerned Scientists (www.ucsusa.org) and NASA.
Threats to Biodiversity 125

curring earlier in the spring (Chen et al. 2011; Willis et al. 2008). Increasing
temperatures are also associated with summer drought and rising mortal-
ity rates of trees in Europe and with devastating outbreaks of tree-killing
beetles in the Rocky Mountains (Carnicer et al. 2011). Because the implica-
tions of global climate change are so far-reaching, biological communities,
ecosystem functions, and climate need to be carefully monitored over the
coming decades (Nelson et al. 2013). Global climate change will also have
an enormous effect on human populations in coastal areas affected by ris-
ing sea level and increased hurricane impacts, as well as in areas that are
already experiencing drought stress and desertification. In many areas
of the world, crop yields will decline because of less favorable growing
conditions (Hannah et al. 2013). Crop yields are predicted to decline by an
average of 8% in tropical areas of Africa and South Asia where chronic,
severe hunger is already an enormous problem, and by 30% or more in cer-
tain populous countries such as Brazil and Indonesia. The disadvantaged
people of the world will be least able to adjust to these changes and will
suffer the consequences disproportionately. However, all countries of the
world will be affected, and it is time for people and their governments to
recognize the urgent need to address global climate change.
It is likely that, as the climate changes, many existing protected areas
will no longer preserve the rare and endangered species that currently
live in them (Hole et al. 2011). We need to establish new conservation areas
now to protect sites that will be suitable for these species in the future,
such as sites with large elevational and latitudinal gradients. Potential
future migration routes, such as north–south river valleys, also need to
be identified and established now (Nuñez et al. 2013). If species are in
danger of going extinct in the wild because of global climate change, the
last remaining individuals may have to be maintained in captivity. An-
other strategy that we need to consider is to move isolated populations
of rare and endangered species to new localities at higher elevations and
closer to the poles, where they can survive and thrive. This approach has
been termed assisted colonization. There is considerable debate within the
conservation community about whether assisted colonization represents
a valid strategy or whether it is too problematic because of the potential
for transplanted species to become invasive species in their new ranges.
Even if global climate change is not as severe as predicted, establishing
new protected areas can only help to protect biodiversity.
Although the prospect of global climate change is cause for great con-
cern, it should not divert our attention from the massive habitat destruction
that is the principal current cause of species extinctions. Many conservation
biologists believe that preserving intact ecosystems, restoring degraded
ones, and increasing the connectivity of existing protected areas are the
most important and immediate priorities for conservation, especially in
the marine environment. These protected areas will also facilitate the mi-
gration of species as they adjust their ranges in response to a changing
climate. Regardless, it is imperative that we reduce our use of fossil fuels
126 Chapter 4

and protect and replant forests in order to decrease levels of greenhouse


gases (Kintisch 2013).

Overexploitation
People have always hunted and harvested the food and other resources
they need to survive. As long as human populations were small and their
methods of collection unsophisticated, people could sustainably harvest
and hunt the plants and animals in their environment. As human popula-
tions have increased, however, our use of the environment has escalated,
and our methods of harvesting have become dramatically more efficient. In
many areas, this has led to an almost complete depletion of large animals
from many biological communities and the creation of strangely “empty”
habitats (Redford 1992).
Technological advances mean that, even in the developing world, guns
are used instead of blowpipes, spears, or arrows for hunting in tropical
forests and savannas. Small-scale local fishermen now have outboard mo-
tors on their boats, which allow them to harvest wider areas more rapidly.
Powerful motorized fishing boats and enormous factory ships harvest fish
from the world’s oceans and sell them on the global market (Figure 4.18).
Even in preindustrial societies, intense exploitation, particularly for meat,

FIGURE 4.18 Intensive harvesting has reached crisis levels in many of the
world’s fisheries. These bluefin tuna (Thunnus thynnus) are being transferred
from a fishing trawler to a factory ship, aboard which huge quantities of fish are
efficiently processed for human consumption. Such efficiency can result in mas-
sive overfishing. (Photograph © Images & Stories/Alamy.)
Threats to Biodiversity 127

has led to the decline and extinction of local species of birds, mammals,
and reptiles (Doughty 2013).
Some traditional societies imposed restrictions on themselves to prevent
overexploitation of jointly owned common property or natural re-
sources (Cinner and Aswani 2007). For example, the rights to spe-
cific harvesting territories were rigidly controlled, and hunting Today’s vast human
and harvesting in certain areas were banned. There were often population and improved
prohibitions against harvesting female, juvenile, and undersized
technology have resulted
animals. Certain seasons of the year and times of the day were
in unsustainable harvest
closed for harvesting. Certain efficient methods of harvesting
levels of many biological
were not allowed. (Interestingly enough, these restrictions, which
resources.
allowed some traditional societies to exploit communal resources
on a long-term, sustainable basis, are almost identical to some of
the fishing restrictions regulators have imposed on or proposed
for many fisheries in industrialized nations.)
Such self-imposed restrictions on using common-property resources
are often less effective today. In much of the world, resources are exploited
opportunistically without restraint. In economic terms, a regulated com-
mon-property resource sometimes becomes an open-access resource and
available to everyone without regulation. The lack of restraint applies to
both ends of the economic scale—the poor and hungry as well as the rich
and greedy. If a market exists for a product, local people will search their
environment to find and sell it. Sometimes traditional groups will sell the
rights to a resource, such as a forest or mining area, for cash. In rural areas,
the traditional controls that regulate the extraction of natural products
have generally weakened. Whole villages are mobilized to systematically
remove every usable animal and plant from an area of forest. Where there
has been substantial human migration, civil unrest, or war, controls of
any type may no longer exist. In countries beset with civil conflict, such
as Somalia, Cambodia, the former Yugoslavia, the Democratic Republic of
the Congo, and Afghanistan, firearms have come into the hands of rural
people. The breakdown of food distribution networks in countries such
as these leaves the resources of the natural environment vulnerable to
whoever can exploit them (Loucks et al. 2009).
In some areas, populations of large primates, such as gorillas and chim-
panzees, as well as ungulates and other mammals may be reduced by
80% or more by hunting. Populations of certain species may be eliminated
altogether, especially those that occur within a few kilometers of a road
(Lindsey et al. 2013). In many places, hunters are extracting animals at a
rate six or more times greater than the resource base can sustain. The result
is a forest with a mostly intact plant community that is lacking its animal
community (Galetti and Dirzo 2013). Without large animals, many plant
species lack effective seed dispersal and decline in abundance.
The decline in animal populations caused by intensive hunting for
food, which has been termed the bushmeat crisis, is a major concern for
wildlife officials and conservation biologists, especially throughout Africa
(www.bushmeat.org; Linder and Oates 2011). Furthermore, eating primate
128 Chapter 4

bushmeat increases the possibility that new diseases will be transmitted


to human populations; both HIV and the Ebola virus have been linked
to such practices (Calmy et al. 2015). Solutions involve restricting the sale
and transport of bushmeat, restricting the sale of firearms and ammuni-
tion, closing roads following logging, extending legal protection to key
endangered species, establishing protected reserves where hunting is not
allowed, and most importantly, providing alternative protein sources to
reduce the demand for bushmeat (Moro et al. 2013).

International wildlife trade


International trade in wildlife is valued at over $240 billion per year, and
trade within developing countries is rapidly increasing as well (Goss and
Cumming 2013; Hinsley et al. 2015). Major exporters are primarily located
in the developing world, often in the tropics. Most major importers are in
the developed countries and East Asia; they include Canada, China, the
European Union, Japan, Singapore, Taiwan, and the United States.
The legal and illegal trade in wildlife is responsible for the decline of
many species (Nijman et al. 2011). For example, in 2015, 1338 African rhinos
were killed—the highest number in one year since the current poaching
epidemic began in 2008 (IUCN 2016). Most of these killings were motivated
by the international market for rhino horn, which is believed by some to
have medicinal properties. Overharvesting of butterflies by insect col-
lectors; of parrots by aviculturists; of orchids, cacti, and other plants by
horticulturists; of marine molluscs by shell collectors (see Figure 5.5); and
of tropical fishes by aquarium hobbyists are further examples of the target-
ing of whole biological communities to supply an enormous international
demand (Table 4.3; Raghavan et al. 2013).
It has been estimated that 1 billion tropical fish are sold annually world-
wide for the aquarium market, and many times that number are killed
during collection and shipping (Whittington and Chong 2007). The inter-
national trade in other live animals is similarly large: 1 million reptiles and
250,000 birds per year. More than 400 endangered bird species are hunted
or traded (Regueira and Bernard 2012). Or consider primates: as many as
70,000 live primates are exported each year (Nijman et al. 2011); China is
the largest exporter of live primates and the United States is the largest
importer (primarily for research). In an attempt to regulate and restrict this
trade, many declining species are listed as protected under the Conven-
tion on International Trade in Endangered Species (CITES; see Chapter 6).
Listing species with CITES has often protected species or groups of species
from further exploitation.
A striking example of overexploitation is the worldwide trade in frog
legs; each year Indonesia exports the legs of roughly 100 million frogs to
western European countries for luxury meals. There is no information on
how this intensive harvesting affects frog populations, forest ecology, or ag-
riculture, and perhaps not surprisingly, the names of the frog species on the
shipping labels are often wrong, which adds to the difficulty in quantifying
the extent of the problem (Warkentin et al. 2009; www.traffic.org). Adding
Threats to Biodiversity 129

Table 4.3 Major Groups Targeted by the Worldwide Trade in Wildlife


Total individuals
traded in 2013
Group (alive or dead) Comments
Primates 637,000 Mostly used for biomedical research; also for pets, zoos,
circuses, and private collections.
Birds 3.3 million Zoos and pets. Mostly perching birds, but also legal and
illegal trade of about 80,000 parrots.
Reptiles 15.6 million Zoos, pets, and raw skins. Also 10–15 million raw skins.
Reptiles are used in some 50 million manufactured products.
Mainly come from the wild, but increasingly from farms.
Ornamental fish 3.7 million Most saltwater tropical fish come from wild reefs and may be
caught by illegal methods that damage other wildlife and
the surrounding coral reef.
Reef corals 2 million Reefs are being destructively mined to provide aquarium
decor and coral jewelry.
Orchids 9–10 million 70% of the species listd by CITES are orchids.Approximately
10% of the international trade comes from the wild.
Cacti 7–8 million Approximately 15% of traded cacti come from the wild;
smuggling is a major problem.
Sources: Data from CITES Trade Database, Appendix 2 for 2013 (as compiled in National Geographic 2015); WRI
2005; Karesh 2005; and Nijman et al. 2011.

to the stress placed on frog populations is the use of some species, such as
the Lake Titicaca frog (Telmatobius coleus), as medicine (Figure 4.19). This
critically endangered frog is endemic to Lake Titicaca, which straddles Peru
and Bolivia (De la Riva and Reichle 2014). Harvesting of
these frogs is believed to have decreased the population
by 80%, but vigorous conservation and education efforts
seem to be helping (Reading et al. 2011).
Yet another example is the enormous demand for
sea horses (Hippocampus spp.) in China. The Chinese
use dried sea horses in their traditional medicine be-
cause they resemble dragons and are believed to have
a variety of healing powers. About 54 tons of sea horses
are consumed in China per year—roughly 19 million

FIGURE 4.19 The Lake Titicaca frog (Telmatobius


coleus) has been harvested to the point of extinction from
this South American lake for use as an exotic dish and as
a treatment for asthma and other ailments. The frogs are
blended whole and then sold in Peru and Bolivia as drinks,
frog extract (as shown here), canned food, and flour. Also
added into the product is maca (Lepidium meyenii), a rad-
ish-like plant from the Andes that used to be endangered
but is now grown commercially. (Photograph courtesy of
Walter Silva [ATFSS-Lima].)
130 Chapter 4

animals. Sea horse populations throughout the world are being decimated
to supply this ever-increasing demand, with the result that international
trade in sea horses is increasingly monitored and regulated by interna-
tional treaty (Vincent et al. 2014).
Although most international trade in wildlife is legal and therefore can
be regulated, an estimated $10 billion per year is not. A black market links
poor local people, corrupt customs officials, rogue dealers and criminal
gangs, and wealthy buyers who do not question the sources from which
they buy. This trade has many of the same characteristics, the same prac-
tices, and sometimes the same players as the illegal trade in drugs and
weapons, and it is extremely widespread and highly profitable. Confront-
ing those who perpetuate such illegal activities has become a major and
dangerous job for international law enforcement agencies.

Commercial harvesting
Governments and industries often claim that they can avoid the
Species can often recover overharvesting of wild species by applying modern scientific
when they are protected management. As part of this approach, an extensive body of
from overexploitation. literature has developed in wildlife and fisheries management
and in forestry to describe the maximum sustainable yield: the
greatest amount of a resource, that can be harvested each year
and replaced through population growth without detriment to
the population. In many real-world situations, however, industry repre-
sentatives and government officials managing commercial harvesting op-
erations may lack the key biological information that is needed to make
accurate calculations. Not surprisingly, attempts to harvest at high levels
can lead to abrupt species declines.
For many marine species, direct exploitation is less important than the
indirect effects of commercial fishing (Burgess et al. 2013). Many marine
vertebrates and invertebrates are caught incidentally during fishing opera-
tions; most of these organisms, referred to as bycatch, are killed or injured
in the process. Between 25% and 75% of the catch in fishing operations is
dumped back into the sea. The declines of skates, rays, and seabirds of 148
species have all been linked to their wholesale death as bycatch (Zydelis et
al. 2009). The huge number of sea turtles and dolphins killed by commercial
fishing boats as bycatch resulted in a massive public outcry and led to the
development of improved nets to reduce these accidental catches. The de-
velopment of improved nets and hooks, as well as other methods to reduce
bycatch, is an active area of current fisheries research (Riskas et al. 2016).
One of the most heated debates over the harvesting of wild marine
species has involved the hunting of whales. The debate is due in part to
the strong emotional attachment to whales that many people in Western
countries have (see Figure 3.14). After recognizing that many whale spe-
cies had been hunted to dangerously low levels, the International Whaling
Commission finally banned all commercial whaling in 1986. Despite that
ban, certain species remain at densities far below their original numbers.
Threats to Biodiversity 131

Those species include the blue whale (Balaenoptera musculus) and the northern
right whale (Eubalaena glacialis), which have been protected since 1967 and
1935, respectively. The densities of other species, such as the gray whale
(Eschrichtius robustus), appear to have recovered, however (Table 4.4). The
slow recovery of some species may be due to continued hunting, both legal
and illegal. Whale hunting by the Japanese fleet continues under the dubi-
ous claim that additional scientific data are needed to assess the status of
whale populations.
Finding the best methods to protect and manage the remaining indi-
viduals in overharvested populations is a priority for conservation biolo-
gists. As we will see in Chapter 8, projects linking the conservation of bio-
diversity to local economic development represent one possible approach.
In some cases, this linkage may be made possible by acknowledging the
sustainable harvesting of a natural resource with a special certification
that allows producers to receive a higher price for their product. Certi-
fied timber products and seafoods are already entering the market, but it

Table 4.4 Worldwide Populations of Whale Species Harvested by


Humans
Numbers
prior to Present Primary diet
Species whalinga numbers items IUCN statusb
Baleen whales
Blue 200,000 10,000– Plankton Endangered
25,000
Bowhead 56,000 25,500 Plankton Least concern
Fin 475,000 60,000 Plankton, fish Endangered
Gray (Pacific stock) 23,000 15,000– Crustaceans Least concern
22,000
Humpback 150,000 60,000 Plankton, fish Least concern
Minke 140,000 1,000,000 Plankton, fish Least concern
Northern right Unknown 1,300 Plankton Endangered
Sei 250,000 54,000 Plankton, fish Endangered
Southern right 100,000 10,000 Plankton Least concern
Toothed whales
Beluga Unknown 200,000 Fish, squid, Near threatened
crustaceans
Narwhal Unknown 50,000 Fish, squid, Near threatened
crustaceans
Sperm 1,100,000 360,000 Fish, squid Vulnerable
Sources: American Cetacean Society (www.acsonline.org); IUCN Red List.
a
Preexploitation population numbers are highly speculative; genetic evidence suggests the popula-
tions might have been even greater (Roman and Palumbi 2003; Alter et al. 2007).
b
Status is determined by a combination of numbers, threats, and trends. For example, numbers of
Southern right whales are low but are increasing.
132 Chapter 4

remains to be seen whether they will have a significant positive


Certifying timber, sea- effect on biodiversity, particularly among increasingly affluent
food, and other products consumers in China and other Asian countries. It is also un-
as sustainable may known whether the regulations associated with certification
be a way to prevent will be enforced in practice and will restrain the ongoing threats
overharvesting. (Christian et al. 2013).
National parks, nature reserves, marine sanctuaries, and
other protected areas can also be established to conserve over-
harvested species. In some cases, cooperative actions involving
international organizations, individual countries, and nongovernmental
conservation organizations are needed to prevent overharvesting (Oster-
blom and Bodin 2012). When harvesting can be reduced or stopped by the
enforcement of international regulations such as CITES and comparable
national regulations, species may be able to recover. Elephants, sea otters,
sea turtles, seals, and certain whale species provide hopeful examples
of species that have recovered—once overexploitation was stopped—in
certain places in the world (Lotze et al. 2011; Magera et al. 2013).

Invasive Species
Species that become established and proliferate in new (i.e., nonhistorical)
ranges where they cause environmental harm are considered invasive
(Mack et al. 2000). Not unlike chemical pollution and other risk factors
mentioned above, invasive species have negative effects on ecosystems
because the native organisms have not evolved adaptations to deal with
the new conditions they impose. In most cases, the conditions imposed
by an invasive species are novel because it has only recently been intro-
duced to that ecosystem. However, native species can also become inva-
sive, usually in response to human alterations to their ecosystem, such as
the availability of human food waste supporting great increases in corvids
(crows and ravens) and their predation on other bird’s nests in conifer-
ous forests (Carey et al. 2012). Invasive species represent threats to 42%
of the endangered species in the United States and have had particularly
severe impacts on bird and plant species (Pimentel et al. 2005). The effects
of invasive species have been estimated to cost countries from 1.4% up
to 12% of their GDP, amounting to billions of dollars per year (Marbuah
et al. 2014). Globally, over half of all recent animal extinctions are attrib-
utable in whole or in part to the effects of invasive species (Clavero and
García-Berthou 2005).
Species invasions have occurred by a variety of means:
• European colonization. European settlers arriving at new colonies re-
leased hundreds of European bird and mammal species into places like
New Zealand, Australia, North America, and South Africa to make the
countryside seem familiar and to provide game for hunting. Numer-
Threats to Biodiversity 133

ous species of fish (trout, bass, carp, etc.) have been widely
released to provide food and recreation. In some cases, food Invasive species may
sources introduced to support game species have disrupt- displace native species
ed food webs and led to declines in populations of game through competition for
species. limiting resources, prey
• Agriculture, horticulture, aquaculture. Large numbers of plant on native species to the
species have been introduced and grown as crops, orna- point of extinction, or
mentals, pasture grasses, or soil stabilizers. Many of these alter the habitat so that
species have escaped from cultivation and have become natives are no longer able
established in local ecosystems. As aquaculture develops, to persist.
there is a constant danger of more plant species escaping
and becoming invasive in marine and freshwater environ-
ments (Xu et al. 2014).
• Accidental transport. Species are often transported unintentionally
(Hulme 2015). For example, weed seeds are accidentally harvested
with commercial seeds and sown in new localities; rats, snakes, and
insects stow away aboard ships and airplanes; and disease-causing
microbes, parasitic organisms, and insects travel along with their host
species, particularly in the leaves and roots of plants and the soil of
potted plants (Liebhold et al. 2012). Around 70% of the nonnative for-
est pest insects in the United States arrived on imported living plants.
Seeds, insects, and microorganisms on shoes, clothing, and luggage
can be transported across the world in a few days by people travel-
ing by plane. Ships frequently carry organisms in their ballast tanks,
releasing vast numbers of bacteria, viruses, algae, invertebrates, and
small fish into new locations. Large ships may hold up to 150,000 tons
of ballast water. Governments are now developing regulations to re-
duce the transport of species in ballast water, such as requiring ships
to exchange their ballast water 320 km (200 miles) offshore in deep
water before approaching a port (Costello et al. 2007).
• Biological control. When a nonnative species becomes invasive, a com-
mon solution is to release an animal species from its original range
that will consume the pest and hopefully control its numbers. While
biological control can be dramatically successful, there are cases in
which a biological control agent itself has become invasive, attacking
native species along with (or instead of) the intended target species
(Elkinton et al. 2006). For example, an herbivorous weevil (Larinus
planus) introduced into North America to control invasive Eurasian
thistles (Carduus spp.) has been found to attack populations of rare na-
tive North American thistles (Cirsium spp.; Havens et al. 2012; Louda
et al. 1997). In order to minimize the probability of such effects, species
being considered as biological control agents are tested before release
to determine whether they will restrict their feeding to the intended
target species.
134 Chapter 4

Threats posed by invasive species


The negative effects of invasive species can be both direct and indirect.
Abundant research has documented these for many types of organisms
and may cascade through an ecosystem. When invasive plant species domi-
nate a community, the diversity and abundance of native plant species and
the insects that feed on them may show a corresponding decline (Heleno
et al. 2009). Evidence also indicates that invasive plants can even reduce
the diversity of soil microbe species (Callaway et al. 2004). Mechanisms
whereby invasive species cause harm include the following:
• Competition for resources. Through their dramatic expansion, invasive
organisms may use up or block access to resources needed by native
organisms, including food, space, light, and even mates. Nonnative
plants can compete with native plants for pollinators or overwhelm
them with nonnative pollen (Beans and Roach 2015). Introduced Eu-
ropean earthworm species are currently outcompeting native species
in soil communities across North America, with negative effects on
certain ground-nesting songbirds as well as potentially enormous, but
as yet unknown, consequences for underground biological communi-
ties (Loss and Blair 2011).
• Predation and parasitism. Extinction risk from invasion is often associat-
ed with predation. Introduced rats are a primary management concern
on many islands; when they were successfully controlled in areas of
New Zealand, many populations of native insects (Ruscoe et al. 2013)
and birds (O’Donnell and Hoare 2012) showed dramatic increases.
American bullfrogs (Lithobates catesbeianus) are voracious predators
and invasive in many parts of the world; a study in Argentina found
them to consume 40 different prey taxa, including several species of
native frogs and crustaceans (Quiroga et al. 2015).
• Changes in ecosystem processes. Both cheatgrass (Bromus tectorum) and
tamarisk (Tamarix spp.), two of the most pervasive weeds in the west-
ern United States, indirectly kill native plants by providing fuel that
promotes wildfires (Drus 2013). Soil microbial communities are also
different under tamarisk trees, lacking important symbionts for native
plants (Meinhardt and Gehring 2013). The presence of domesticated
cats has been shown to affect a wide range of ecological processes,
including bird migration patterns, seed dispersal, and breeding and
parental behaviors (Medina et al. 2014).
• Alteration of abiotic conditions. Some invasive plants have been found to
alter soil properties, including moisture, acidity, and enzymes, which
can facilitate more invasions or otherwise harm native species (Kueb-
bing et al. 2014). Both invasive plants and animals can alter the struc-
tural environment (Gutiérrez et al. 2014): animals through behaviors
such as burrowing and plants because of their particular growth forms,
which may differ in shape or density from those of natives. Such chang-
es to the environment may then affect ecosystem processes.
Threats to Biodiversity 135

Invasive species may have more than one of these effects on an eco-
system at once, particularly those that are in the middle of the food chain,
as insects and other invertebrates usually are. Insects introduced both
deliberately, such as European honeybees (Apis mellifera), and accidentally,
such as fire ants (Solenopsis invicta) and gypsy moths (Lymantria dispar),
can build up huge populations, both competing with native animals and
preying on native plants and animals. At some localities in the southern
United States, the diversity of insect species declined by 40% following the
invasion of nonnative fire ants, and there was a similarly large decline in
native birds (Figure 4.20).

Invasive species on oceanic islands


The isolation of oceanic island habitats encourages the evolution of a
unique assemblage of endemic species (see Chapter 7), but it also leaves
those species particularly vulnerable to depredations by invading species.
Many plants that grow on islands with few herbivores do not produce the
bad-tasting, tough vegetative tissue that discourages herbivores, nor do
they have the ability to resprout rapidly following damage. Some island
birds that have evolved in the absence of predators have lost the power
of flight and simply build their nests on the ground. Such species often

0.040
Bobwhite abundance (birds/observer hour)

0.030
Bobwhite abundance
declines by 75%

0.020

0.010

0 4 8 12 16 20
Years after arrival of fire ants

FIGURE 4.20 The abundance of northern bobwhites (Colinas virginianus) in


Texas has been declining over a 20-year period following the arrival of the non-
native red fire ant (Solenopsis invicta). The fire ants may directly attack and dis-
turb bobwhites, particularly at the nestling stage, and may compete with them
for food items, such as insects. (After Allen et al. 1995; bobwhite photograph
courtesy of Steve Maslowski/US Fish and Wildlife Service; fire ant photograph
courtesy of Richard Nowitz/USDA ARS.)
136 Chapter 4

succumb rapidly when invasive herbivores or predators are introduced.


Mammals and other vertebrates introduced to islands often prey efficiently
on endemic animal species and have grazed some native plant species to
extinction (Garzón-Machado et al. 2010). Moreover, island species often
have no natural immunity to mainland diseases. When nonnative domesti-
cated species (e.g., chickens, ducks) arrive, they frequently carry pathogens
or parasites that, though relatively harmless to the carriers, devastate the
native populations (e.g., wild birds).
The introduction of just one nonnative species to an island may cause
the local extinction of numerous native species. The brown tree snake (Boiga
irregularis) has been introduced onto a number of Pacific islands, where it
is devastating endemic bird populations. The snake eats eggs, nestlings,
and adult birds. On Guam alone, the brown tree snake has driven 10 of 13
forest bird species extinct (Perry and Vice 2009). The government spends
$3 million per year on attempts to control the brown tree snake population,
including aerial drops of dead mice injected with acetaminophen (a poison
for snakes), so far without success.
A quarter of the native bird species of New Zealand have gone extinct
since the arrival of humans on the islands and the resulting introduction
of many invasive species to which the birds (and other native organisms)
were not adapted (Russell et al. 2015a). New Zealand’s national symbol,
the kiwi (Apteryx spp.), is an endemic flightless bird that did not evolve
with mammalian predators such as dogs (Canis lupus familiaris) and stoats
(Mustela ermine, a type of weasel). It is estimated that New Zealand is
losing 2% of the kiwi population each year to these predators. In 2014,
the Predator Free New Zealand program was launched with the goal of
extirpating all invasive mammals from the country, beginning with its
smallest islands.

Invasive species in aquatic habitats


The diversity and abundance of both freshwater and saltwater species is
frequently found to decline in association with aquatic invaders, which
include plants, fishes, and invertebrates (Gallardo et al. 2016). Whereas
invasions in freshwater environments are often more readily noticed, a
widespread survey found that there are 329 invasive marine species and
that 84% of marine areas worldwide are affected by at least one of them
(Molnar et al. 2008). Freshwater ecosystems are somewhat similar to oce-
anic islands in that they are isolated habitats and are thus at increased risk
of invasion. There has been a long history of introductions of nonnative
commercial and sport fishes into lakes, such as the introduction of the Nile
perch (Lates niloticus) into Lake Victoria in East Africa, which was followed
by the extinction of numerous endemic cichlid fishes. Often the introduced
nonnative fishes are larger and more aggressive than the native fish fauna,
and they may eventually drive the local fish to extinction. But once these
invasive species are removed from aquatic habitats, the native species are
sometimes able to recover (Vredenburg 2004).
Threats to Biodiversity 137

Aggressive aquatic invaders also include plants and invertebrates. One


of the most alarming invasions in North America was the arrival in 1988 of
the zebra mussel (Dreissena polymorpha) in the Great Lakes (Strayer 2009).
This small, striped native of the Caspian Sea apparently was a stowaway
in the ballast tanks of a European cargo ship (Figure 4.21). Within 2
years, zebra mussels had reached almost unbelievable densities of 700,000
individuals per square meter in parts of Lake Erie, encrusting every hard
surface and choking out native mussel species in the process. As they
continue to spread throughout the waters of the United States, these non-
native molluscs are causing enormous economic damage, estimated at $1
billion per year, to fisheries, dams, power plants, water treatment facilities,
and boats, as well as devastating the aquatic communities they encounter
(Pimentel et al. 2005). Both the zebra mussel and, more recently, the quagga
mussel (Dreissena rostriformis bugensis) have become sig-
nificant problems not only in the United States, but in
Europe as well (Figure 4.22).
One-third of the worst invasive species in aquatic
environments are aquarium and ornamental species,
which are traded worldwide to the tune of $25 billion
per year (Keller and Lodge 2007). Notable examples
include water hyacinth (Eichornia crassipes), an orna-
mental plant that creates huge floating mats (Figure
4.23). These mats deprive submerged plants of light;
use up oxygen in the water, thereby killing fish and
turtles; are barriers to animals that fish; and even act
as vectors of disease by creating habitat for mosquitoes
that carry malaria and for a snail that hosts a flatworm
that causes disease in humans. Another such invader,
Caulerpa taxifolia, is a species of marine algae that has
blanketed the seafloor, covering native algae, sea grass,
and sessile animals, in the Mediterranean Sea. Both
water hyacinth and Caulerpa are listed by the Interna-
tional Union for the Conservation of Nature (IUCN) as
among the world’s top 100 worst invasive species.

The ability of species to become invasive


The great majority of introduced species do not survive
outside of their native ranges, and of those that do, only
FIGURE 4.21 This current meter was
a small fraction (perhaps less than 1%) are capable of in-
retrieved from Lake Michigan after a crust
creasing and spreading in their new locations. Why are
of thousands of zebra mussels made it in-
certain species able to invade and dominate new habi- operable. Such encroachment is typical of
tats and displace native species so easily? These species the tiny molluscs, which also encrust and
may be better suited to taking advantage of disturbed destroy native mussel species and other
conditions than native species (Dukes et al. 2011), par- organisms. (Photograph by M. McCormick,
ticularly if the disturbance is new to that system and courtesy of NOAA/Great Lakes Environ-
alters the way physical or chemical resources are made mental Research Laboratory.)
138 Chapter 4

2006 2007

Netherlands
North Sea

Amsterdam

Germany

Belgium Location of
first record

2009 2012

FIGURE 4.22 The rapid spread


of the Eastern European quagga mus-
sel (Dreissena rostriformis bugensis)
in the Netherlands since it was first
detected in the Rhine–Meuse estuary
in 2006. In Europe, the quagga mussel
has traveled between 23 and 383 km/y
(mean 120 ±53.8 km/y). (After Mat-
thews et al. 2014.)

available (Sher and Hyatt 1999). Human activity that causes disturbances
may create unusual environmental conditions, such as higher or lower min-
eral nutrient levels, increased or decreased incidence of fire, or enhanced or
lowered light availability, to which nonnative species are sometimes better
adapted than native species. In fact, the highest concentrations of invasive
species are often found in those habitats that have been most altered by
human activity. Many of the threats to biodiversity mentioned above can
Introduction to Conservation Biology 1E Primack/Sher
Sinauer Associates cause or exacerbate invasion; for example, ocean acidification was found to
Morales Studio increase predation by invasive snails on native oysters by 48% (Sanford et
Primack_Sher1E_04.22 Dateal. 01-04-16 1/15/16habitats are altered by global climate change, they become
2014). When
even more vulnerable to invasion (Bradley et al. 2012).
As mentioned above, native species can become invasive within their
home ranges, usually when they are suited to the ways in which humans
have altered the environment, and may be almost as much of a concern
as nonnative invasive species (Carey et al. 2012). Within North America,
Threats to Biodiversity 139

FIGURE 4.23 Water hyacinth (Eichhornia crassipes), which is native to South


America, has become a significant ecological and economic problem in many
other parts of the world, including North America, Africa, and Asia. (© Stephanie
Jackson/Alamy Stock Photo.)

fragmentation of forests, suburban development, and easy


access to garbage have allowed the numbers of coyotes, red
foxes, and certain gull species to increase. Native jellyfishes
have increased in abundance in the Gulf of Mexico because Invasive species often
they use oil rigs and human structures for spawning and feed thrive where human
on plankton blooms stimulated by nitrogen pollution (Duarte activities have changed
et al. 2013). As these aggressive species increase, they do so at the environment.
the expense of other local native species, such as the juvenile
stages of commercially harvested fish, representing a further
challenge to the management of vulnerable native species and
protected areas.
Another explanation for why some introduced species become invasive
is the predator release hypothesis, which attributes their rapid proliferation
in their new habitat to the absence of the specialized natural predators and
parasites that would otherwise control the invaders’ population growth
(Davis 2009). In Australia, for example, introduced rabbits spread uncon-
trollably, grazing native plants to the point of extinction, because there
were no effective checks on their numbers. Australian control efforts have
focused in part on introducing specific diseases as a biological control for
rabbit populations.
140 Chapter 4

In one of the key generalizations of this field, the species that are most
likely to become invasive and have significant effects in a new location
are those species that have already been shown to do so somewhere else
(Ricciardi 2003). A special class of invasive species is made up of those
introduced species that have close relatives in the native biota (Laikre
et al. 2010). When invasive species hybridize with the native species and
varieties, unique genotypes may be eliminated from local populations and
taxonomic boundaries may become obscured (see Chapter 2)—a process
called genetic swamping. This appears to be the fate of native trout species
when confronted by introduced species. In the southwestern United States,
the Apache trout (Oncorhynchus apache) has had its range reduced by habitat
destruction and competition with introduced species. The species has also
hybridized extensively with rainbow trout (O. mykiss), an introduced sport
fish, blurring its identity as a distinct species.

Control of invasive species


Invasive species are considered to be the most serious threat facing the
biota of the US national park system. While the effects of habitat deg-
radation, fragmentation, and pollution can potentially be corrected and
reversed in a matter of years or decades as long as the original species are
present, well-established invasive species may be impossible to remove
from ecosystems. They may have built up such large numbers and become
so widely dispersed and so thoroughly integrated into an ecosystem that
eliminating them may be extraordinarily difficult and expensive (Rinella
et al. 2009).
The threats posed by invasive species are so severe that reducing the
rate of their introduction needs to become a greater priority for conserva-
tion efforts (Liebhold et al. 2012). Governments must pass and enforce
laws and customs restrictions prohibiting the transport and introduc-
tion of nonnative species. In some cases, this may require restrictions
and inspections related to the movement of soil, wood, plants, animals,
and other items across international borders and even through
checkpoints within countries. Currently, vast sums are spent
Governments must act to controlling widespread outbreaks of invasives, but inexpen-
prevent the introduction sive, prompt control and eradication efforts at the time of first
of new invasive species, sighting can stop a species from getting established in the first
to monitor the arrival place (McConnachie et al. 2012; Van Wilgen et al. 2012). Training
and spread of invasives, citizens and protected areas staff to monitor vulnerable habitats
and to eradicate new for the appearance of known invasive species and promptly
populations of invasives. implementing intensive control efforts can be an effective way
to stop the establishment and early spread of a potential new
invader. This strategy may require a cooperative effort on the
part of multiple levels of government and private landowners.
A variety of strategies for controlling and eliminating invasive species
exists, including three general approaches to their removal: pesticides (i.e.,
chemical poisons), mechanical culling (physical removal, such as pulling
weeds and trapping or shooting animals), and biological control (use of
Threats to Biodiversity 141

one organism to control another). Each of these approaches, used alone


or in combination with the others, has both benefits and limitations. In
many cases, use of pesticides is the most effective means of controlling
undesirable species, but the chemicals can be expensive and may hurt
nontarget, desirable species. Mechanical approaches usually have the few-
est nontarget effects but are the most labor-intensive. Biological control,
such as use of a specialist insect or pathogen, can be the cheapest approach
to managing invaders and is usually the most gentle on the ecosystem
as a whole, but biological controls will not usually eradicate the species
in question and may take time to show any effect. An extensive public
education program is often necessary so that people are aware of why
invasive species need to be removed or killed, especially when they are
charismatic mammals such as cats, dogs, and rabbits (Oppel et al. 2011).
An emphasis on “compassion for the ecosystem” is one approach that
has been suggested to help explain why lethal controls are sometimes
warranted (Russell et al. 2015b).
Habitat manipulation, changing agricultural practices, and the use of
crop strains that are more resistant to pests are other methods frequently
employed to prevent invasions or lessen their effects. Using a combina-
tion of these techniques for the long-term management of pests is called
integrated pest management (IPM).
Even though the effects of invasive species are generally considered to
be negative, they may provide some benefits as well, especially when the
habitat is so degraded that the original species are unlikely to reestablish
themselves, even once the invasive species are removed. Invasive plant
species can sometimes stabilize eroding lands, provide nectar for native
insects, and supply nesting sites for birds and mammals (Bateman et al.
2013; Shackelford et al. 2013). In such situations, the trade-offs need to be
evaluated to determine whether the potential benefits of removal will out-
weigh the overall costs. Increasingly, biologists are recognizing the value of
accepting “novel ecosystems” (see Chapter 10) in which a mixture of native
and nonnative species is best suited to the new conditions created by human
activities (Hobbs et al. 2013).

GMOs and conservation biology


A special topic of concern for conservation biologists is the increasing use
of genetically modified organisms (GMOs) in agriculture, forestry, aqua-
culture, and toxic waste cleanup. GMOs are organisms to whose genetic
code scientists have added genes from a different (“source”) species using
the techniques of recombinant DNA technology. Such gene transfers can
be done not only across species, but across taxonomic domains, as when a
bacterial gene that produces a chemical toxic to insects is transferred into
a crop species such as corn. Enormous amounts of cropland—especially
in the United States, Argentina, China, and Canada—have already been
planted with GMOs, mainly soybeans, corn (maize), cotton, and oilseed
rape (canola). Genetically modified animals are under development, with
salmon and pigs showing commercial potential.
142 Chapter 4

Humans have been genetically modifying domesticated crop and ani-


mal species since the dawn of civilization by means of selective breeding,
hybridization, and other forms of artificial selection. However, many spe-
cies being investigated as potential sources of transferable genes, including
viruses, bacteria, insects, fungi, and shellfish, have not previously been
used in breeding programs. Fear of “crossovers” between unrelated species
has resulted in some governments implementing special controls on this
type of research and its commercial applications. There is concern among
some people, especially in Europe, that genetically modified crop species
will hybridize with related species, leading to invasion by new, aggressive
weeds and virulent diseases (Bagla 2010). Additionally, the use of GMOs
could potentially harm noncrop species, such as insects, birds, and soil
organisms that live in or near agricultural fields. Further, some people
want assurances that eating food from GMO crops will not harm their
health or cause unusual allergic reactions.
It is clear that GMOs have the ability to increase crop production to
feed a growing human population, to produce new and cheaper medicines,
and to reduce the use of pesticides on agricultural fields and the pesticide
pollution associated with such use (Lu et al. 2012). (On the other hand, one
hugely popular GMO—the Roundup Ready soybean—has been genetically
engineered by the manufacturer of the herbicide glyphosate, commonly
sold as the weed killer Roundup, so that the crop can be treated with
more—not less—Roundup.)
In summary, the benefits of GMOs need to be examined and weighed
against the unknown potential risks. The best approach involves proceeding
cautiously, investigating GMOs thoroughly before commercial releases are
authorized, and monitoring environmental and health effects after release.

Disease
The increased transmission of disease as a result of human activities is
a major threat to many endangered species and ecosystems. Pathogens
(disease-causing organisms) such as bacteria, viruses, fungi, and protists
can have major effects on vulnerable species and even the structure of
entire ecosystems. Human activities can lead to increased popu-
lations of many pathogens, leading to outbreaks of animal and
human disease. In addition, interaction with humans and their
Increased incidence
domesticated animals exposes wild animals to diseases never
of infectious disease
previously encountered that can reduce the size and density of
threatens wild and
wild populations (Jones et al. 2008; Figure 4.24).
domesticated species as
Disease may be the single greatest threat to some rare species.
well as humans. Transfer The decline of numerous frog populations in visually pristine
of disease between montane habitats across the world is apparently due to the in-
species is a subject of troduction of a nonnative fungal disease. The last population
special concern. of black-footed ferrets (Mustela nigripes) known to exist on its
own in the wild was destroyed by the canine distemper virus in
Threats to Biodiversity 143

Habitat fragmentation and degradation;


high local densities

Wild ducks and


other waterfowl

Agricultural advance Human advance


into wildlife areas into wildlife areas

Chickens Humans

Higher animal Increased travel; higher


densities in modern human densities in urban
agriculture Technology and industrial agriculture areas and villages

FIGURE 4.24 Infectious diseases—such as rabies, Lyme disease, influenza,


bird flu, hantavirus, and canine distemper—spread among wildlife populations,
domesticated animals, and humans as a result of increasing population densities
and the advance of agriculture and human settlements into wild areas. The dia-
gram illustrates the infection and transmission routes of bird flu, which is caused
by a virus that infects wild waterfowl, chickens, and humans. The shaded areas
of overlap indicate that the disease can be shared among the three groups.
Green arrows indicate factors contributing to higher rates of infection; blue
arrows indicate factors contributing to the spread of disease among the three
groups. (After Daszak et al. 2000.)

1987, though a few healthy individuals were caught for a captive breeding
program. One of the main challenges of managing the captive breeding
program has been protecting the captive ferrets from canine distemper,
human viruses, and other diseases; this is being done through rigorous
quarantine measures and subdivision of the captive colony into geographi-
cally separate groups.
White-nose syndrome is a fungal disease that is currently killing mil-
Introduction to Conservation Biology 1E Primack/Sher
lions
SinauerofAssociates
bats across the eastern United States (Thogmartin et al. 2013).
In someStudio/SA
Morales caves, 90% of the bats have died. The disease is characterized
Primack_Sher1E_04.24
by a powdery white fuzz Date on a12-17-15 1/15/16
bat’s snout and other membranous areas
144 Chapter 4

FIGURE 4.25 A little brown


bat (Myotis lucifugus) infected
with white-nose syndrome is
being examined by a research-
er. This fungal disease, which
first appeared in the United
States in 2006, affects a variety
of bat species and has killed
up to 90% of the bats in many
caves. (Photograph courtesy
of Ryan Von Linden/New York
Department of Environmental
Conservation.)

(Figure 4.25). Bats die when the fungus causes skin irritation and the
bat wakes from hibernation in midwinter, when its diet of flying insects
is not available, instead of in spring, depleting its energy reserves and
subsequently starving to death. Discovered in one cave in New York State
in 2006, the disease has spread rapidly across the region, probably during
bat migration. It is possible that cave explorers or bat researchers acciden-
tally introduced the fungus to the United States as a contaminant on their
clothes, boots, or equipment following a visit to a European bat cave. At
this point, the only effective way to prevent its spread to new colonies is
to close bat caves to all human visitors except for scientists who sterilize
their clothes and equipment before entering.
Three basic principles of epidemiology have obvious practical implica-
tions for limiting disease in the management of endangered species. First, a
high rate of contact between host and pathogen or parasite is one factor that
encourages the spread of disease. In general, as host population density
increases, so does risk of disease. In addition, a high density of the infective
stages of a parasite in the environment of the host population can lead to
increased incidence of disease. In natural situations, the rate of infection is
typically reduced when animals migrate away from their droppings, saliva,
old skin, dead conspecifics, and other sources of infection. However, in
unnaturally confined situations, such as habitat fragments, zoos, or even
nature reserves, the animals remain in contact with these potential sources
of infection, and disease transmission increases. Furthermore, at higher
densities, animals have abnormally frequent contact with one another,
and once one animal becomes infected, the parasite can rapidly spread
throughout the entire population.
Second, indirect effects of habitat destruction can increase an organ-
ism’s susceptibility to disease. When a host population is crowded into a
smaller area because of habitat destruction, its members often face lowered
habitat quality and food availability, lowered nutritional status, and less
Threats to Biodiversity 145

resistance to infection in addition to higher contact rates (Zylberberg et al.


2013). Furthermore, crowding can lead to social stress within a population,
which also lowers the animals’ resistance to disease. Pollution may make
individuals more susceptible to infection by pathogens, particularly in
aquatic environments (Harvell et al. 2004). It has even been observed that
biodiversity regulates disease by diluting the number of suitable host spe-
cies or by constraining the sizes of host populations through predation and
competition (Ostfeld 2009). For example, the increased incidence of Lyme
disease and other tick-borne pathogens in some areas has been linked not
only to the local abundance of certain host rodent species, as mentioned
above, but also to the overall loss of local species diversity (Figure 4.26).
Third, in many conservation areas, zoos, national parks, and newly
cleared agricultural areas, species come into contact with other species that
they would rarely or never encounter in the wild—including humans and
domesticated animals—so infections such as rabies, influenza, distemper,
hantavirus, and bird flu can spread from one species to another. Infectious
diseases can spread between wildlife populations, domesticated animals,
and humans as a result of increasing human population densities and the
advance of agriculture and human settlements into wildlife areas. The

(A) (B)

(C)
FIGURE 4.26 (A) The white-footed mouse
(Peromyscus leucopus), one of the main hosts for
the bacterium that causes Lyme disease, increases
in abundance in habitat fragments created by sub-
urban development. (B) Field biologists sampling
mice for the presence of infectious diseases, such
as plague. (C) A black-legged tick, which can be
up to 3 mm (0.12 in.) long, can transfer the Lyme
disease bacterium to a human after acquiring it
from an infected animal host. (A, © Rob and Ann
Simpson/Visuals Unlimited, Inc.; B, from Crowl et
al. 2008; C, courtesy of Michael L. Levin/CDC.)
146 Chapter 4

human immunodeficiency virus (HIV) and the deadly Ebola


Steps must be taken virus both appear to have spread from wildlife populations to
to prevent the spread humans and to domesticated animals. Such examples are likely
of disease in captive to become more common as a result of anthropogenic changes
animals and to ensure to the environment, the increase in international travel, and
that new diseases are not globalization of the economy.
accidentally introduced In zoos, colonies of animals are often caged together in small
into wild populations. areas, and similar species are often housed close to one another.
Consequently, if one animal becomes infected, the pathogen can
spread rapidly to other animals and to related species. Once
they are infected with an exotic disease, captive animals cannot
be returned to the wild without threatening the entire wild population.
Furthermore, a species that is both common and fairly resistant to a disease
can act as a reservoir for the disease, which can then infect populations
of susceptible species. For example, apparently healthy African elephants
can transmit a fatal herpesvirus to related Asian elephants when they are
kept together in zoos. Diseases can spread very rapidly between captive
species kept in crowded conditions.

A Concluding Remark
This chapter has described seven major categories of threats faced by spe-
cies and ecosystems. A study of 181 threatened and endangered species
(Lawler et al. 2002) found that more than 85% of them faced at least four
types of threats, the most common of which were related to habitat destruc-
tion and degradation, interactions with invasive species, and overexploita-
tion. When the threats to biodiversity are well understood, protection and
recovery efforts have the best chance for success.

Summary
„„The major threats to biodiversity are tat is both reduced and divided into two
habitat destruction, fragmentation, deg- or more fragments. Habitat fragmenta-
radation (which includes pollution), cli- tion can lead to the rapid loss of some of
mate change, overexploitation, invasive the remaining species because it creates
species, and disease. All of these threats barriers to the normal processes of dis-
result from the use of the world’s natu- persal, colonization, and foraging. Par-
ral resources by an increasing human ticular fragments may contain altered en-
population. vironmental conditions that make them
„„Habitat destruction threatens rain for- less suitable for the original species.
ests, wetlands, coral reefs, and other „„Environmental pollution eliminates
species-rich communities. many species from ecosystems even
„„Habitat fragmentation is the process where the structure of the community is
whereby a large, continuous area of habi- not obviously disturbed. Environmental
Threats to Biodiversity 147

pollution results in pesticide biomagni- „„Humans have deliberately and acci-


fication; contamination of water with in- dentally moved thousands of species to
dustrial wastes, sewage, and fertilizers; new regions of the world. Some of these
and air pollution resulting in acid rain, nonnative species have become invasive,
excess nitrogen deposition, photochemi- greatly increasing their numbers at the
cal smog, and high ozone levels. expense of native species.
„„Global climate change, including warmer „„Levels of disease often increase when
temperatures and changing precipitation animals are confined to nature reserves,
patterns, is already occurring because of zoos, or habitat fragments and can-
the large amounts of carbon dioxide and not disperse over wide areas. In zoos
other greenhouse gases produced by the and botanical gardens, diseases some-
burning of fossil fuels and deforestation. times spread between related species of
Predicted temperature increases may be animals and plants. Diseases may also
so rapid in coming decades that many spread between domesticated species
species will be unable to adjust their and wild species, and even between hu-
ranges and will become extinct. mans and both wild and domesticated
„„Overexploitation is driving many spe- animals.
cies to extinction and can consequently „„Species may be threatened by a combi-
undermine entire ecosystems. Over- nation of factors, all of which must be
exploitation is caused by increasingly addressed in a comprehensive conserva-
efficient methods of harvesting and mar- tion plan.
keting, increasing demand for products,
and increased access to remote areas.

For Discussion
1. Human population growth is often live or go to school. Why have some
blamed for the loss of biological diver- habitats been preserved and others frag-
sity. Is this valid? What other factors are mented and degraded?
responsible, and how do we weigh their 3. Learn about one endangered species in
relative importance? Is it possible to find detail. What is the full range of immedi-
a balance between providing for increas- ate threats to this species? How do these
ing numbers of people and protecting immediate threats connect to larger so-
biodiversity? cial, economic, political, and legal issues?
2. Consider the most damaged and the
most pristine habitats near where you

Suggested Readings
Bussière, E., L. G. Underhill, and R. Altwegg. 2015. Patterns of bird migration
phenology in South Africa suggest northern hemisphere climate as the
most consistent driver of change. Global Change Biology 21: 2179–2190. The
warming climate in breeding grounds has been found to have impacts on
the timing of bird movements.
148 Chapter 4

Clavero, M. and E. García-Berthou, E. 2005. Invasive species are a leading


cause of animal extinctions. Trends in Ecology and Evolution, 20(3), 110. A
frequently cited documentation of the impact of invasive animals.
Gallardo, B., M. Clavero, M. I. Sánchez, and M. Vilá. 2016. Global ecological
impacts of invasive species in aquatic ecosystems. Global Change Biol-
ogy. 22: 151–163. A review and analysis of 733 evaluations of the effects of
aquatic invaders.
Gibson, L., A. J. Lynam, C. J. Bradshaw, F. He, D. P. Bickford, D. S Woodruff,
and 2 others. 2013. Near-complete extinction of native small mammal
fauna 25 years after forest fragmentation. Science 341: 1508–1510. Forest
fragments gradually lose many of their original species.
Gutiérrez, J. L. and C. Bernstein. 2014. Ecosystem impacts of invasive species.
BIOLIEF 2011. 2nd World Conference on Biological Invasion and Ecosys-
tem Functioning, Mar del Plata, Argentina, 21–24 November 2011. Acta
Oecologica 54 :1–138.
Intergovernmental Panel on Climate Change (IPCC). 2013. Climate Change
2013: The Physical Science Basis. Contribution of Working Group I to the Fifth
Assessment Report of the Intergovernmental Panel on Climate Change. T. F.
Stocker and 9 others (eds). Cambridge University Press, Cambridge, UK.
Comprehensive presentation of the evidence for global climate change,
along with predictions for the coming decades.
Larson, C. 2016. Shell trade pushes giant clams to the brink. Science 351: 323–
324. Tridacna gigas faces extinction from over harvesting.
Lovich, J. E., C. B. Yackulic, J. Freilich, M. Agha, M. Austin, K. P. Meyer, and 4
others. 2014. Climatic variation and tortoise survival: Has a desert species
met its match? Biological Conservation 169: 214–224. Prolonged drought as-
sociated with climate change is impacting endangered desert species.
Primack, R. B. and A. J. Miller-Rushing. 2012. Uncovering, collecting, and
analyzing records to investigate the ecological impacts of climate change:
A template from Thoreau’s Concord. BioScience 62: 170–181. We can see the
effects of climate change happening all around us if we find the right tools
and baseline studies.
Robertson, G., C. Moreno, J. A. Arata, S. G. Candy, K. Lawton, J. Valencia,
and 4 others. 2014. Black-browed albatross numbers in Chile increase in
response to reduced mortality in fisheries. Biological Conservation 169: 319–
333. Seabird numbers increased after the Chilean fishing industry switched
to a system of fast-sinking baited hooks that were less accessible to birds.
Siraj, A. S., M. Santos-Vega, M. J. Bouma, D. Yadeta, D. R. Carrascal, and M.
Pascual. 2014. Altitudinal changes in malaria incidence in highlands of
Ethiopia and Colombia. Science 343: 1154–1158. Malaria spreads to higher
altitudes in warmer years and is predicted to have a wider impact with a
warming climate.
Wilson, M. C., X. Y. Chen, R. T. Corlett, R. K. Didham, et al. 2016. Habitat frag-
mentation and biodiversity conservation: key findings and future chal-
lenges. Landscape Ecology 31(2): 219–227. Many of the impacts of habitat
fragmentation are indirect.
5 Extinction Is Forever
The Meaning of “Extinct” 154 Vulnerability to Extinction 166
Measuring Extinction 160 Problems of Small
Populations 172

The Tasmanian tiger-wolf (Thylacinus cynocephalus) was a marsupial


that lived in Australia and on the island of Tasmania. After competition
with introduced dogs led to its extinction on the mainland, the remaining
populations in Tasmania were destroyed by a combination of habitat loss,
hunting, and disease. By 1933, the species was extinct in the wild, and the
last known individual died in captivity in 1936.
T
he diversity of species found on the Earth
has been increasing since life first originated.
The increase has not been steady; rather,
it has been characterized by periods of high rates of
speciation followed by periods of minimal change and
episodes of mass extinction (Raup and Sepkoski 1982;
Ward 2004). In addition to this overall increase, there
have been five episodes of mass extinction in the fos-
sil record, occurring at intervals ranging from 60 mil-
lion to 155 million years (Figure 5.1). These episodes,
which ended the Ordovician, Devonian, Permian, Tri-
assic, and Cretaceous periods, could be called natural
mass extinctions. The most famous is the extinction of
the dinosaurs during the Late Cretaceous, 65 million
years ago, after which mammals achieved dominance
in terrestrial communities. The most massive extinc-
tion took place at the end of the Permian, 250 million
years ago, when about 95% of all marine animal spe-
cies and half of all animal families are estimated to
have gone extinct (Wake and Vredenburg 2008). As
David Raup (1988) observed, “If these estimates are
even reasonably accurate, global biology (for higher
organisms at least) had an extremely close brush with
total destruction.” It is quite likely that some massive
disturbance—such as widespread volcanic eruptions,
a collision with an asteroid, or both—caused the dra-
matic change in the Earth’s climate that resulted in
152 Chapter 5

Geologic Millions of Groups experiencing


period years ago mass extinction
Extinction
Quaternary 0.01 Current: many groups; extinctions
largely the result of human activities
Tertiary
Extinction
65 Cretaceous: reptiles (dinosaurs);
many marine species, including
Cretaceous many foraminiferans and molluscs

Jurassic Extinction
180 Triassic: 35% of animal families,
Triassic including many reptiles and marine molluscs
Extinction
250 Permian: 50% of all animal families,
Permian including over 95% of marine species;
many trees, amphibians, most bryozoans
and brachiopods, all trilobites
Carboniferous
Extinction
345 Devonian: 30% of animal families,
Devonian including agnathan and placoderm
fishes and many trilobites
Silurian

Ordovician
Extinction
500 Ordovician: 50% of animal families,
Cambrian including many trilobites
Bar width represents relative
number of living groups

Figure 5.1 Relative numbers of animal families over geologic time. Al-
though the total number of species groups on Earth has increased over the eons,
a large percentage of these groups disappeared during each of five episodes of
natural mass extinction (named in boldface at left). The most dramatic period of
loss occurred about 250 million years ago, at the end of the Permian period. A
sixth mass extinction episode (red arrow at top of figure) began during the pres-
ent geologic period and will continue for decades to come.

the end of so many species. Another speculation is that there might have
been a massive release of methane gas from beneath the ocean floor—a
“big burp,” if you will. Such an event not only would have released toxic
plumes but almost certainly would have affected the climate, since meth-
ane is an even more potent greenhouse gas than carbon dioxide. It took
about 80–100 million years of evolution for Earth’s biota to regain the num-
ber of families lost during the Permian extinction.
Species go extinct even in the absence of violent disturbance. One spe-
cies may outcompete another for a vital resource, or predators may drive
prey species to extinction. Extinction is as much a natural process as spe-
ciation is. If extinction is a natural process, however, why is the current
loss of species of such concern? The answer lies in the relative rates of
extinction and speciation as well as in the causes of extinction. Speciation
Introduction to Conservation Biology 1E aPrimack/Sher
is typically slow process, occurring through the gradual accumulation
Sinauer Associates
Morales Studio/SA
Primack_Sher1E_05.01 Date 1-11-16
Extinction Is Forever 153

of mutations and shifts in allele frequencies over thousands, if


not millions, of years. As long as the rate of speciation equals or The current rate of
exceeds the rate of extinction, biodiversity will remain constant species loss is unprece-
or increase. In past geologic periods, the loss of existing species dented, unique, and
was eventually balanced and then exceeded by the evolution of irreversible. Ninety-
new species. However, current extinction rates are much greater nine percent of current
than speciation rates, and more than 99% of modern species ex- extinctions are caused
tinctions are linked to human activity (Pimm and Jenkins 2005; in some way by human
Wake and Vrendenberg 2008). We are presently in the midst of activities.
a sixth extinction episode, caused by human activities rather
than a natural disaster (Ceballos et al. 2015).
Extinction is a top concern of conservation biology because
the number of threatened species is so large. About 21% of the world’s
remaining bird species are threatened with extinction. Table 5.1 shows
certain animal groups for which the danger is particularly severe, includ-
ing turtles, manatees, rhinoceroses, and penguins. The danger is even

Table 5.1 Numbers of Species Threatened with Extinction in Major


Groups of Animals and Plants
Number of species Percentage of
Approximate threatened with species threatened
Group number of species extinction with extinction
Vertebrates
Fishes 28,000 2,084 7a
Amphibians 6,338 2,303 36
Reptiles 9,400 1,018 11a
Crocodiles 23 10 43
Turtles 228 170 75
Birds 10,052 2,097 21
Penguins 18 11 61
Mammals 5,499 1,462 27
Primates 418 227 54
Manatees, dugongs 5 4 80
Horses, tapirs, rhinos 16 14 88
Plants
Gymnosperms 18,000 521 3a
Angiosperms 260,000 9,614 4a
(flowering plants)
Palms 361 275 82
Fungi 100,000 3 0
Source: Data from IUCN 2011 (www.iucnredlist.org). Data include the categories critically endangered,
endangered, vulnerable, and near threatened.
a
Low percentages reflect inadequate data due to the small number of species evaluated. For example,
11% of reptiles are listed as endangered, but only about one-third of species have been evaluated. Among
reptile species that have been evaluated, 31% are considered endangered.
154 Chapter 5

greater for other large groups: 27% of mammal species and 36% of am-
phibian species are threatened (IUCN 2013). Some other groups of species,
such as fishes (9%) and reptiles (12%), face lower threat levels. Some plant
species are also at risk; palms are especially vulnerable. Most groups of
plants, fungi, fishes, and insects are not well known, and for these groups
the extinction risk cannot be accurately determined. Extinction rates are
likely to be much higher once threats faced by poorly known groups of
species are assessed (Hoffmann et al. 2010; McClenachan et al. 2012).

The Meaning of “Extinct”


The word extinct has many nuances, and its meaning can vary somewhat
depending on the context.
• A species is considered extinct when, after a thorough search, no mem-
ber of the species is found alive anywhere in the world: the Tasmanian
tiger-wolf (Thylacinus cynocephalus), for example, is extinct (see the
chapter opening photo).
• If individuals of a species remain alive only in captivity or in other
human-controlled situations, the species is said to be extinct in the
wild: The St. Helena ebony tree (Trochetiopsis ebenus) (Figure 5.2) is
nearly extinct in the wild, although it grows well under cultivation.
• A species is locally extinct, or extirpated, when it is no longer found in
a specific area it once inhabited, but is still found elsewhere in the wild:
“The purple bladderwort can no longer be found in Fairhaven Bay, but
it can still be found in other nearby locations.” This is in contrast to a
species that no longer lives anywhere in the world, sometimes referred
to as being globally extinct.

Figure 5.2 The St. Helena ebony tree


(Trochetiopsis ebenus) is endemic to the
island of St. Helena in the South Atlantic
Ocean. The wild population has been re-
duced to just two individuals on the side of
a cliff. Because of these low numbers, this
species will almost certainly go extinct in
the wild. The species will remain alive in
cultivation. (© fotoFlora/Alamy.)
Extinction Is Forever 155

• Some conservation biologists speak of a species as being functionally


extinct (or ecologically extinct) if it persists at such reduced numbers
that its effects on the other species in its community are negligible:
“Tigers are functionally extinct because so few remain in the wild that
their effect on prey populations is insignificant.” Species that are func-
tionally extinct are generally at great risk of becoming globally extinct.
In addition to the various nuances of the term extinct, conservation biolo-
gists work with a variety of categories, including endangered, vulnerable,
and threatened, that more specifically describe the status of species. These
categories are discussed in detail in Chapter 6.

The current, human-caused mass extinction


The global diversity of species reached an all-time high in the present
geologic period. Many groups of organisms—such as insects, vertebrates,
and flowering plants—reached their greatest diversity about 30,000 years
ago. Since that time, however, species richness has slowly decreased as
one species—Homo sapiens—has asserted its dominance. In our need to
consume natural resources, humans have increasingly altered terrestrial
and aquatic environments at the expense of other species.
The first noticeable effects of human activity on extinction rates can be
seen in the elimination of large mammals from Australia and North and
South America at the time humans first colonized these continents tens of
thousands of years ago. Shortly after humans arrived, approximately 80%
of the megafauna—mammals weighing more than 44 kg (100 pounds)—be-
came extinct. These extinctions were probably caused directly by hunting
and indirectly by the burning and clearing of forests and grasslands and
the introduction of invasive species and new diseases. On all continents,
paleontologists and archaeologists have found an extensive record of pre-
historic human alteration and destruction of habitat coinciding with high
rates of species extinctions (Sandom et al. 2014). For example, deliberate
burning of savannas, presumably to encourage plant growth for browsing
wildlife and thereby to improve hunting, has been occurring for 50,000
years in Africa.
Extinction rates during the last 2000 years are best known for terrestrial
vertebrates, especially birds and mammals, because these species are con-
spicuous and are therefore well studied. Extinction rates for the other 99%
of the world’s species are just rough guesses at present because scientists
have not systematically searched the remote sites where many of these spe-
cies occur to determine whether they still exist. For example, a species of
giant tortoise (Chelonoidis elephantopus) known from only one island in the
Galápagos was thought to have gone extinct shortly after Charles Darwin
visited in 1835, but it was recently rediscovered on another island and its
identity confirmed using DNA sequencing methods.
On the other hand, some species presumed to be extant (still living) may
actually be extinct; the Yangtze river dolphin (Lipotes vexillifer), for example,
has not officially been designated as extinct, but no individuals could be
156 Chapter 5

found during an intensive survey in 2006 (Fisher and Blomberg 2012), and
many of the previous sightings are probably not reliable (Turvey 2008).
Sometimes it is difficult to determine whether a species is truly extinct.
In 2004, for example, ornithologists in North America announced sight-
ings of an ivory-billed woodpecker (Campephilus principalis) in an Arkansas
swamp forest—decades after this bird was believed to have gone extinct.
Since then, however, intensive efforts to find and conclusively identify
existing individuals of the species have been unsuccessful (Gotelli et al.
2012; Scott et al. 2008; Solow et al. 2012).
One set of estimates based on the best available evidence indicates that
79 species of mammals have gone extinct and another 126 are critically
endangered—a total 3.7% of all known species (Pimm et al. 2014). Amphib-
ians have fared far worse (8.1%). The majority of human-caused extinctions
have occurred in the last 150 years. The extinction rate for birds was about
0–5 species every 25 years during the period from 1500 to 1725, but it rose
to 8–12 species every 25 years during the period from 1750 to 1850. After
1850, the extinction rate rose again to more than 16 species every 25 years.
This increase in the rate of extinction reflects the growing intensity of the
threat to biodiversity.
One trend to note is that all of the earliest extinctions docu-
mented by humans were on islands. In fact, 89% of known bird
Island species have had extinctions since 1500 have been on islands, even though greater
higher rates of extinction than 80% of bird species occur on continents (Butchart et al.
than mainland species. 2006). It has been determined from archeological findings that
Freshwater species 20% of bird species worldwide have gone extinct since the ar-
are more vulnerable to rival of humans on Pacific islands 30,000 years ago, due to both
extinction than marine hunting and the introduction of new predators (Steadman 1995).
species. In contrast, extinctions of birds in mainland areas were first
observed about 1800, and they have been increasing since then
(Figure 5.3). In the future, mainland species will account for
an increasing proportion of extinctions. Studying island extinc-
tions has taught us much about population dynamics that is applicable to
mainland conservation efforts.
Fortunately, extinction rates appear to have declined since 1950, in
part due to deliberate efforts to protect rare species in danger of going
extinct. These numbers can be misleading, however, because of the cur-
rent practice of not declaring a species extinct until decades after any
individuals of the species have been found. In coming years, numerous
species will be declared to have gone extinct during the past half century.
In the last decade, a number of species that were not found despite in-
tensive searches were declared extinct, including the Monteverde golden
toad of Costa Rica (Bufo periglenes) and the Alaotra grebe (Tachybaptus
rufolavatus) of Madagascar. Many species not yet listed as extinct—and
some species that have never been documented at all—have been reduced
to such low numbers by human activities that their ability to persist is
Extinction Is Forever 157

18
Island bird species
16
Mainland bird species

14
Extinctions per 25 years

12

10

0
1500 1550 1600 1650 1675 1700 1725 1750 1775 1800 1825 1850 1875 1900 1925 1950 1975
Year

FIGURE 5.3 Rates of extinctions of bird species during 25-year intervals since
1500. Initially, extinctions were almost exclusively of island species, but extinc-
tions of mainland species have increased since 1800. (After Butchart et al. 2006.)

uncertain. In Britain, the country with the best-kept biologi-


cal records, extinction rates for major groups of species have Many species today
been in the range of 1%–5% per century, but are predicted to are represented only by
increase in the present century (Hambler et al. 2011). scattered populations
A count of extant species may also mask the true scale of the with reduced numbers
current extinction crisis. For many species, a few individuals in
of individuals. Although
scattered small populations might persist for years, decades,
these isolated popul-
or centuries (isolated individuals of woody plants, in particu-
ations could persist for
lar, can persist for hundreds of years), but their ultimate fate is
years or decades, the
global extinction, either because they are functionally extinct or
ultimate fate of such
because negative environmental effects have yet to be fully felt.
Remaining individuals of species that are doomed to extinction species is extinction.
have been called “the living dead” or “committed to extinction”
(Hanski et al. 1996; Janzen 2001). There are certainly many spe-
cies in this category in the remaining fragments of species-rich
tropical forests. The presumed eventual loss of species following habitat
Introduction to Conservation Biology 1E Primack/Sher
destruction
Sinauer and fragmentation is called the extinction debt (Tilman et al.
Associates
1994; Dullinger
Morales Studio/SA et al. 2013). For example, it is estimated that 80% of forest-
Primack_Sher1E_05.03
dependent vertebrate species Date 1-11-16 2/22/16
will disappear because of destruction that
has already occurred (Wearn et al. 2012).
158 Chapter 5

FIGURE 5.4 This tropical


checkered-skipper butterfly
(Pyrgus oileus) is one of the
remaining extant species on
Barro Colorado Island, Panama.
Fully 23 known species of but-
terflies have gone locally extinct
there since the 1930s, although
the actual number is probably
much higher, given how many
cryptic butterfly species exist.
(© Rick and Nora Bowers/Ala-
my Stock Photo.)

Local extinctions
In addition to the global extinctions that are a primary focus of con-
servation biology, many species are experiencing a series of local
extinctions, or extirpations, across their range (Leidner and Neel 2011).
Where habitats are degraded and destroyed, populations of plants and
animals go extinct. For example, 23 species of butterflies have gone locally
extinct on Barro Colorado Island, Panama, since the 1930s, representing
20% of the common butterfly species there (Basset et al. 2015) (Figure 5.4).
Many carnivores have experienced local extinctions in recent history across
the globe (Ripple et al. 2014). Biological communities are impoverished by
such local extinctions.
Concord, Massachusetts, was first assessed for wildflower species in
the 1850s by the famous naturalist and philosopher Henry David Thoreau.
Twenty-seven percent of the native species seen by Thoreau and other
nineteenth-century Concord botanists could not be found when the area
was surveyed 150 years later (Miller-Rushing and Primack 2012; Primack
et al. 2009). A further 36% of the species now persist in only one or two
populations and are therefore vulnerable to extinction. In some cases, only
a few individual plants remain of species that were formerly common.
Certain groups, such as orchids and lilies, have shown particularly severe
losses. A combination of forest succession, invasive species, air and water
pollution, grazing by deer, habitat destruction and fragmentation, and
now climate change has contributed to these species losses in Concord. In
other examples, according to surveys by state Natural Heritage programs,
4%–8% of the plant species formerly found in Hawaii, New York, and Penn-
sylvania can no longer be found in these states. Due to deforestation over
the last 200 years, 34% of bird species and 26% of vascular plant species in
Singapore have become locally extinct (Brook et al. 2003). And in a survey
of one part of the Indonesian island of Sumatra, only 3 of 12 populations
Extinction Is Forever 159

of Asian elephants known from the 1980s were still present 20


years later (Hedges et al. 2005). Species-rich tropical
It is estimated that there are an average of 200 populations rain forests are being
per species (Hughes and Roughgarden 2000). While some spe- lost at a rate of 1% a
cies have just a few populations, other species may have thou- year, a rate believed to
sands of populations. The loss of populations is roughly equal result in the destruction
to the proportion of a habitat that is lost, so the world’s popula- of more than 13,500
tions are being lost at a far higher rate than its species. When biological populations
90% of an extensive grassland ecosystem is destroyed, 90% of
each day. Population
the populations of plants, animals, and fungi there will also
losses eventually result in
be lost. Tropical rain forests contain at least half of the world’s
species extinctions.
species, and they are being lost at the rate of about 1% per year.
This represents a loss of 5 million populations per year (1%
of the 500 million tropical forest populations), or about 13,500
populations per day.
These large numbers of local extinctions serve as important biologi-
cal warning signs that something is wrong in the environment. Action is
needed to prevent further local extinctions as well as global extinctions.
The loss of local populations not only represents a loss of biodiversity
but also diminishes the value of an area for nature enjoyment, scientific
research, and the provision of crucial materials to local people in subsis-
tence economies.

Extinction rates in aquatic environments


In contrast with the large amount of information we have on extinct ter-
restrial species, only about 14 species—4 marine mammals, 5 marine birds,
1 fish, and 4 molluscs—are known to have gone extinct in the world’s vast
oceans during historic times (Régnier et al. 2009). This number of extinc-
tions is almost certainly an underestimate, because marine species are not
nearly as well known as terrestrial species, but it also may reflect a greater
resiliency of marine species to disturbance. Regardless, the significance
of these losses may be greater than the numbers suggest. Many marine
mammals are top predators, and their loss could have major effects on
marine communities. Some marine species are the sole species of their
genus, family, or even order, so the extinction of even a few of them could
represent a serious loss to global biodiversity.
The oceans were once considered so enormous that it seemed unlikely
that marine species could go extinct; many people still share this view-
point. However, as marine coastal waters become more polluted and spe-
cies are harvested more intensely, even the vast oceans will not provide
safety from extinction (McCauley et al. 2015) (Figure 5.5). Many species
of whales and large fishes have declined by 90% or more because of over-
harvesting and other human activities and are in danger of extinction.
Freshwater species appear to have a higher extinction rate than marine
species. The modern extinction rate for North American freshwater fishes
is conservatively estimated to be 877 times greater than the background
extinction rate (Burkhead 2012). The fishes of California are particularly
160 Chapter 5

Figure 5.5 (A) The endangered (A)


white abalone (Haliotis sorenseni), a
type of mollusc, in its natural habi-
tat on the coast of California. Several
abalone species have reached danger-
ously low numbers due to overharvest-
ing by humans, both for their muscular
foot, which is a delicacy, and for their
beautiful shells, from which jewelry is
made (B). Shown here is an antique
pin in the shape of the now extinct
Huia bird from New Zealand. Abalo-
nes are also threatened by withering
shell syndrome, caused by the bacte-
rium Candidatus xenohaliotis californi-
ensis, for which a cure has yet to be
found. Due to their low numbers, some
scientists consider them functionally
extinct (Haw 2013). (A, Courtesy of
NOAA; B, © Marc Tielemans/Alamy (B)
Stock Photo.)

vulnerable because of the state’s scarcity of water and its intense develop-
ment: 6% of California’s 129 native fish species are already extinct, and
51% are in danger of extinction. Large numbers of fishes and aquatic in-
vertebrates, such as molluscs, are in danger of extinction because of dams,
pollution, irrigation projects, overharvesting, invasive species, disease, and
general habitat damage.

Measuring Extinction
Quantifying rates of extinction and likelihood of extinction is critical for
understanding the magnitude of the problem and how best to address it.
Extinction Is Forever 161

Calculating extinction rates requires information about many facets of


organisms and their environments, including the extent to which humans
are playing a role.

Background extinction rates


To better understand how calamitous the present extinction rates are, we
can compare them with the natural extinction rates that would prevail
regardless of human activity. What is the natural rate of extinction in the
absence of human influence? Natural background extinction rates can be
estimated by looking at the fossil record. On average, an individual species
lasts about 1 million–10 million years before it goes extinct or evolves into
a new species (Mace et al. 2005; Pimm and Jenkins 2005). Since there are
perhaps 10 million species on the Earth today, we can predict that 1–10 of
the world’s species would be lost per year as a result of a natural extinction
rate of 0.0001%–0.00001% per year. These estimates are derived from stud-
ies of wide-ranging marine animals, so they may be lower than natural
extinction rates for species with narrower distributions, which are more
vulnerable to habitat disturbance; however, they do appear to be applicable
to terrestrial mammals. The current observed rate of extinction for birds
and mammals is 1% per century (or 0.01% per year), which is 100–1000
times greater than would be predicted based on background rates of ex-
tinction. Putting it another way, about 100 species of birds and mammals
were observed to go extinct between 1850 and 1950, but the natural rate of
extinction would have predicted that, at most, only 1 species would have
gone extinct. Therefore, the other 99 extinctions can be attributed to the
effects of human activities.

Extinction rate predictions and the island biogeography


model
Studies of island communities have led to general rules on the distribution
of biodiversity, synthesized as the island biogeography model1 by Mac-
Arthur and Wilson (1967). The island biogeography model was built to
explain the species–area relationship: islands with large areas have more
species than do islands with smaller areas (Figure 5.6). This rule makes
intuitive sense because a large island will tend to have a greater variety
of local environments and community types than a small island. In addi-
tion, large islands allow a larger number of separate populations per spe-
cies and larger sizes per individual population, increasing the likelihood
of speciation and decreasing the probability of local extinction of newly
evolved as well as recently arrived species.

1
A scientific model is a physical, conceptual, or mathematical representation of something
observed in nature that can help us understand it and in some cases make predictions.
Models are used to illustrate and explore specific aspects of a real object or phenomenon
and so are necessarily simplifications of reality.
162 Chapter 5

Cuba

Puerto
Rico Redonda
Jamaica

Hispaniola Saba Montserrat

1000
FIGURE 5.6 The number
of species on an island can be
predicted from the island’s area.
In this figure, the number of
Number of species

100 Hispaniola
species of reptiles and amphib-
ians is shown for seven islands Puerto Rico Cuba
in the West Indies. The number Jamaica
of species on large islands such
as Cuba and Hispaniola far ex- 10
ceeds that on the tiny islands of Montserrat
Saba and Redonda. (After Wilson Saba
1989.) Redonda

1
10 100 1000 10,000 100,000
Area of island (km2)

The species–area relationship can be summarized by the empirical


formula
S = CAZ

where S is the number of species on an island, A is the area of the island,


and C and Z are constants. The exponent Z determines the slope of the
relationship between species and area. The values for C and Z depend on
the types of islands being compared (tropical vs. temperate, dry vs. wet,
etc.) and the types of species involved (birds vs. reptiles, etc.). Z values
are typically about 0.25, with a range from 0.15 to 0.35. Island species with
restricted ranges, such as reptiles and amphibians, tend to have Z values
near 0.35, while widespread mainland species tend to have Z values closer
to 0.15. Values of C tend to be high in groups, such as insects, that contain
many species, and they tend to be lower in groups, such as birds, that con-
tain relatively few species (Connor and McCoy 2001).
Imagine the simplest situation, in which C = 1 and Z = 0.25, for raptorial
birds on a hypothetical archipelago:
S = (1)A0.25

Introduction to Conservation Biology 1E Primack/Sher


Sinauer Associates
Extinction Is Forever 163

The formula predicts that islands of 10, 100, 1000, and 10,000 km2 in area
would have 2, 3, 6, and 10 species, respectively. It is important to note that
a tenfold increase in island area does not result in a tenfold increase in the
number of species; with this equation, each tenfold increase in island area
increases the number of species by a factor of approximately 2.
The island biogeography model has been empirically validat-
ed and is accepted by most biologists (Quammen 1996; Triantis The island biogeography
et al. 2008; Chen and He 2009). For numerous groups of plants model can be used
and animals, it has been found to describe the observed species to predict how many
richness reasonably well, explaining about half of the variation species will go extinct
in numbers of species. For example, actual data from three Carib-
due to habitat loss. The
bean islands conform to this relationship: with increasing area,
2 2 2 model can also be used to
Nevis (93 km ), Puerto Rico (8959 km ), and Cuba (114,524 km )
predict how many species
have 2, 10, and 57 species of anolis lizards, respectively; with a
will remain in protected
C of 0.5 and a Z of 0.35, the islands would be predicted to have
areas of different sizes.
2, 12, and 30 species, respectively.
In their classic text, MacArthur and Wilson (1967) also hy-
pothesized that the number of species occurring on an island
represents a dynamic equilibrium between colonization by (and
evolution of) new species and extinctions of existing
species. Starting with an unoccupied island, the number FIGURE 5.7 The island biogeography
of species will increase over time, since more species model describes the relationship between
will be arriving (or evolving) than will be going extinct, the rates of colonization and extinction
until the rates of extinction and immigration are bal- in islands. The immigration rates (blue
anced (Figure 5.7). Species establishment rates will be and red curves) on unoccupied islands
higher for large islands than for small islands because are initially high, as species with good
large islands represent larger targets for dispersing or- dispersal abilities rapidly take advantage of
the available open habitats. The immigra-
tion rates slow as the number of species
increases and sites become occupied. The
extinction rates (green and gold curves)
increase with the number of species on the
island; the more species on an island, the
Island is near
mainland greater the likelihood that a species will
Island go extinct at any time interval. Coloniza-
is small
tion rates will be highest for islands near a
mainland population source, since species
Rate of immigration of new species

Island is can disperse over shorter distances more


far from easily than longer ones. Extinction rates
mainland are highest on small islands, where both
population sizes and habitat diversity are
Island low. The number of species present on
is large an island reaches equilibrium when the
colonization rate equals the extinction rate
Extinction rate

Small, far Large,


near island (circles). The equilibrium number of spe-
island
cies is greatest on large islands near the
mainland and lowest on small islands far
from the mainland. (After McArthur and
0 Number of species Wilson 1967.)
164 Chapter 5

ganisms and are more likely to have open habitat suitable for colonization.
Extinction rates will be lower on large islands than on small islands be-
cause large islands will have greater habitat diversity and greater numbers
of populations. Furthermore, rates of immigration of new species will be
higher for islands near the mainland than for islands farther away, since
mainland species will be able to disperse to near islands more easily than
to distant islands. The model predicts that for any group of organisms, such
as birds or orchids, the number of species found on a large island near a
continent will be greater than that on a small island far from a continent.

Extinction rates and habitat loss


In addition to describing species diversity in natural systems, species–
area relationships have been used to predict the number and percentage
of species that would become extinct if habitats were destroyed (Rompré
et al. 2009). The calculation assumes that reducing the area of natural
habitat on an island would effectively result in a smaller island that
would support fewer species. This model has great utility because it can
be extended to nature reserves or other natural areas on the mainland
that are surrounded by damaged habitat (Chittaro et al. 2010). The re-
serves can be viewed as “habitat islands” in a “sea” of unsuitable habi-
tat. The model predicts that when 50% of an island (or habitat island)
is destroyed, approximately 10% of the species occurring on the island
will be eliminated (Figure 5.8). If these species are endemic to that is-
land—that is, if they occur there and nowhere else—they will become
extinct. When 90% of the habitat is destroyed, 50% of the species will be

100
Percentage of species originally found in a given area

90
FIGURE 5.8 According to the island 50% habitat loss;
biogeography model, the number of 10% species loss
species present in an area increases
asymptotically—that is, it rises sharply
and then levels off, as shown by the red
curve in this example. The shape of the
curve differs from region to region and
among different species groups, but this 50
90% habitat loss;
model gives a general indication of the
50% species loss
relationship between habitat loss and
species loss. Here, if the area of habi-
tat is reduced by 50%, then 10% of the
species in the group will be expected to
disappear; if the habitat is reduced by
90%, half the species will be lost. Stating
this in another way, a system of pro- 10
tected areas covering 10% of a country
could be expected to include 50% of the 10 50 100
country’s species. Percentage of total area preserved
Extinction Is Forever 165

lost; and when 99% of the habitat is gone, about


75% of the original species will be lost. The island
of Singapore can be used as an example. Over the
last 180 years, over 99% of its original forest cover
has been removed; the model estimates that about
70% of its forest species should have been lost. In
fact, between 1923 and 1998, more than 90% of Sin-
gapore’s native birds were lost, with higher rates of
loss for large ground birds and for insectivorous
birds of the forest canopy (Castelletta et al. 2000)
(Figure 5.9).
Predictions of extinction rates based on habitat
loss vary considerably because each species–area
relationship is unique. Using the conservative es-
timate that 1% of the world’s rain forest is being
destroyed each year, Wilson (1989) estimated that Figure 5.9 The banded kingfisher (Lace-
0.2%–0.3% of all rain forest species will be lost per do pulchella) is locally extinct in Singapore.
year. Assuming a total of 5 million rain forest spe- (© PanuRuangjan/Getty Images.)
cies worldwide, that would result in a loss of 10,000
to 15,000 species per year, or 34 species per day.
This estimate predicts that species extinctions by 2050 will be up to 35%
in tropical Africa, 20% in tropical Asia, 15% in the Neotropics, and 8%–10%
elsewhere (MEA 2005). Extinction rates might in fact be higher because the
highest rates of deforestation are occurring in countries with large con-
centrations of rare species, and because large forest areas that remain are
increasingly being fragmented by roads and development projects (Arima
et al. 2015; Walker et al. 2013).
And yet, these predictions of high extinction rates do not seem to be
coming true in the best-studied groups of birds and mammals. For ex-
ample, there are 17 large areas of the world, including the Amazon basin,
the trans-Himalayan region, northern Canada, and the island of New
Guinea, where there have been no recorded extinctions of vertebrates in
the last 200 years (Sanjayan et al. 2012). Extinction rates may be lower than
predicted in these areas because threatened species are being given extra
protection, and perhaps because hotspots particularly rich in endemic
species are being targeted for conservation. Lower extinction rates could
also be observed because species that are committed to extinction are able
to persist at low numbers for several generations. Regardless of which
estimate gives the most reasonable prediction, all estimates indicate that
tens of thousands—if not hundreds of thousands—of species are headed
for extinction within the next 50 years (McCallum 2015; Pimm et al 2015).
The time required for a given species to go extinct following a reduction
in area or fragmentation of its range is a vital question in conservation biol-
ogy, and the island biogeography model makes no prediction about how
long it will take. Small populations of some species may persist for decades
or even centuries in habitat fragments, even though they are committed to
166 Chapter 5

extinction (Sharma et al. 2014). Of the species that will eventually be lost,
the best estimates predict that half will be lost in 50 years from a 1000-ha
fragment, while half will be lost in 100 years from a 10,000-ha fragment
(Brooks et al. 1999). Certain forest mammals in Australia have an expected
persistence time of less than 10 years in small habitat fragments under 10
ha, 50 years in 40–80-ha fragments, and 100 years in 300-ha fragments
(Laurance et al. 2008b). In situations in which there is widespread habitat
destruction followed by recovery, such as in New England and Puerto Rico
over the last several centuries, species may be able to survive in small num-
bers in isolated fragments and then reoccupy adjacent recovering habitat.
Even though 98% of the forests of eastern North America were cut down,
the clearing took place in a patchwork fashion over hundreds of years, so
forest always covered half of the area, providing refuges for mobile animal
species such as birds.
This ability of species to persist for several generations and several
decades in habitat fragments may also be why there have not been more
observed species losses. This conclusion has two important implications
for conservation. First, many species will go extinct in coming decades as
their populations continue to decline in these fragmented habitats. And
second, the persistence of species in fragmented habitats provides a nar-
row window in which conservation actions have the potential to rescue
declining species from extinction, as described in later chapters.

Vulnerability to Extinction
Populations (and ultimately species) that are declining in number are likely
to go extinct unless the cause of decline is identified and corrected (Martin
et al. 2012; Peery et al. 2004). As Charles Darwin pointed out more than 150
years ago in On the Origin of Species (1859):
To admit that species generally become rare before they become extinct,
to feel no surprise at the rarity of the species, and yet to marvel greatly
when the species ceases to exist, is much the same as to admit that
sickness in the individual is the forerunner of death—to feel no surprise
at sickness, but when the sick man dies, to wonder and to suspect that
he died of some deed of violence.

Past decline has been found to be the number one predictor of future
decline (Di Marco et al. 2015). However, once a population is actively
declining, it may be too late to prevent its extinction (see “The extinction
vortex” on p. 188). Thus, conservation biologists have sought to determine
what features of species or populations might be predictive of future
decline or extinction. Some groups of species clearly are more vulnerable
to extinction than others. In Europe, a higher proportion of reptile species
are threatened with extinction than plants, birds, mammals, or fishes
(Dullinger et al. 2013). Ecologists have observed that across taxonomic
Extinction Is Forever 167

lines, there are features that increase a species’ risk of extinction, and
through statistical modeling, they have identified those features that are
most predictive of extinction (Di Marco et al. 2015; Purvis et al. 2000):
• Narrow geographic range. This feature, which defines endemic species,
has been found to be the most predictive of extinction or population
decline (Botts et al. 2013; Di Marco et al. 2015). This conclusion is in-
tuitive: if the whole range is affected by human activity or a natural
disaster, the species may become extinct (Hanna and Cardillo 2013).
Bird species on oceanic islands and fish species confined to a single lake
or watershed are good examples of species with limited ranges that
are especially vulnerable to global climate change. Many tropical bird
species with narrow ranges will face increasing threats of extinction
due to climate change in the coming decades (Şekercioğlu et al. 2012).
Furthermore, a narrow range often (but not always) encompasses only
one or a few populations or a small population size.
• Only one or a few populations. Any one population of a species may be-
come extinct as a result of chance factors, such as earthquake, fire, an
outbreak of disease, or human activity. Species with many populations
are less vulnerable to extinction than are species with only
one or a few populations. This feature is linked to the previ-
ous feature because species with few populations also tend Species most vulnerable
to have narrow geographic ranges (Figure 5.10). to extinction have the
• Small population size. As we will see in the next section, small following characteristics:
populations are more likely to go locally extinct than large narrow geographic
populations because of their greater vulnerability to loss of range, only one or a
genetic diversity and to demographic and environmental few populations, small
variation. Species that characteristically have small popula- populations, declining
tion sizes, such as large predators and extreme specialists, population size, and
are more likely to become extinct than species that typically being hunted or
have large populations (Bulman et al. 2007). At the extreme harvested by people.
are species whose numbers have declined to just a few indi-
viduals. A special category is species with a widely fluctuat-
ing population size in which the population is sometimes
small.
• Island habitat. As mentioned above, the highest species extinction rates
during historic times have occurred on islands. This is not surprising,
given that species on islands often have limited range areas, small
population sizes, and small numbers of populations and are more
likely to be endemic (Régnier et al. 2009). Of the terrestrial animal
and plant species known to have gone extinct from 1600 to the present,
almost half were island species, even though islands represent only a
tiny fraction of the Earth’s land surface. Island species usually have
evolved and undergone speciation with a limited number of competi-
tors, predators, and pathogens, which makes them particularly vulner-
168 Chapter 5

St. Louis, MO

Cincinnati, OH

Providence, RI

Cumberland, OH

Reported historical ABB range


Confirmed ABB state
Captive ABB population
Confirmed ABB county
Attempted ABB reintroduction state
Attempted ABB reintroduction county

FIGURE 5.10 Although the American burying beetle (Nicrophorus america-


nus) was once widespread in the eastern and central United States (outlined in
red), its range is now greatly reduced, which makes it vulnerable to extinction.
It is found in the wild in only two separate areas of the central United States and
on Block Island in Rhode Island. Intensive efforts have been initiated to deter-
mine the cause of this decline and develop a recovery plan. The species is also
being bred in captivity. (After O’Meilla 2004, with updates from the US Fish and
Wildlife Service.)

able to these threats if they are introduced. Species extinction rates peak
soon after humans occupy an island and then decline after the most
vulnerable species are eliminated. In general, the longer an island has
been occupied by people, the greater its percentage of extinct biota. In
Madagascar, 72% of the 9000 plant species are endemic, and 189 spe-
cies are threatened with extinction. The lemurs are also endemic to
Madagascar, and most species of these unique primates are threatened
(Schwitzer et al. 2014) (Figure 5.11).

Introduction to Conservation Biology 1E Primack/Sher


Extinction Is Forever 169

• Hunting or harvesting by people. Overharvesting can rapidly reduce the


population size of a species. If hunting and harvesting are not regu-
lated, either by law or by local customs, the species can be driven to
extinction. Utility has often been the prelude to extinction, as was the
case with the dodo, now a symbol of modern extinctions.
Some additional features of species or populations have been found to
put them at risk:
• Large home range. A species in which individuals or social groups need
to forage over wide areas is prone to die off when part of its range is
damaged or fragmented by human activity.
• Large body size. Large animals tend to have large in-
dividual ranges, have low reproductive rates, re-
quire more food, and be hunted by humans. Top
carnivores, especially, are often killed by humans
because they compete with humans for wild game,
sometimes damage livestock, and are hunted for
sport. Larger animals are generally more vulner-
able to the effects of habitat fragmentation (Riverá
Ortíz et al. 2014). Within groups of species, the larg-
est species is often the most prone to extinction; that
is, the largest carnivore, the largest lemur, the largest
whale will go extinct first.
• Slow reproduction. In some analyses, a low number
of offspring in each reproductive event was one of
the most predictive features of a species becoming
endangered (Purvis et al. 2000). This characteris-
tic often overlaps with large body size: elephants,
whales, giraffes, rhinos, and gorillas all usually have
only one or two young at a time. Other traits associ-
ated with slow reproduction are slow growth rates,
late sexual maturity, long gestation periods, and long
intervals between reproductive events. These traits
mean that a species cannot compensate easily for in-
creased mortality rates (McArthur and Wilson 1967). Figure 5.11 More than 30 species
• Limited dispersal ability. Species unable to adapt phys- of lemurs are all endemic to Madagas-
iologically, genetically, or behaviorally to changing car, a large island off the eastern coast of
environments must either migrate to more suitable Africa. Their endemism, that they live on
an island, their small population sizes and
habitat or face extinction. The rapid pace of anthro-
numbers of populations, and even the fact
pogenic changes often precludes evolutionary adap-
that they are primates (a group with many
tation, leaving migration as a species’ only alterna- endangered species), puts them at higher
tive. Terrestrial species that are unable to cross roads, anticipated risk of extinction. Shown
farmland, or disturbed habitats are doomed to ex- here is a Coquerel’s sifaka and its young
tinction as their original habitat deteriorates. Disper- (Propithecus coquereli). (© Mint Images/
sal is important in the aquatic environment as well, Frans Lanting/Getty Images.)
170 Chapter 5

where dams, point sources of pollution, channelization, and sedimenta-


tion can limit movement. Limited dispersal ability may explain in part
why freshwater fauna such as mussels and snails are more likely to be
extinct or threatened with extinction than are dragonfly species, which
are strong fliers.
• Seasonal migration. Species that migrate seasonally depend on two or
more distinct habitat types. If either one of those habitat types is dam-
aged, the species may be unable to persist. The billion songbirds of 120
species that migrate each year between the northern United States and
the Neotropics depend on suitable habitat in both locations to survive
and breed. In addition, if barriers to dispersal are created by roads,
fences, or dams between the needed habitats, a species may be unable
to complete its life cycle (Wilcove and Wikelski 2008). Salmon species
that are blocked by dams from swimming up rivers and spawning are
striking examples of this problem (Figure 5.12).
• Little genetic variation. Genetic variation within a population can some-
times allow a species to adapt to a changing environment (Forsman
2014); (see the section “Loss of genetic diversity” on p. 172). Species
with little or no genetic variation may have a greater tendency to be-
come extinct when a new disease, a new predator, or some other fac-
tor alters their environment. However, one study determined that for
mammals, genetic variation, while a contributing factor, was not as
important as other, ecological factors, such as body size, for extinction
risk (Polishchuk et al. 2015).
• Specialized niche requirements. Once a habitat has been altered, it may
no longer be suitable for specialized species (Tingley et al. 2013). For

Figure 5.12 Aggrega-


tions of salmon (Oncorhyn-
chus spp.) migrate upstream
to spawning pools. Impedi-
ments to their movement
(such as dams), or anything
else (such as harvesting)
that threatens these ag-
gregations, will severely
threaten populations.
(© Thomas Kline/Design
Pics/Getty Images.)
Extinction Is Forever 171

example, wetland plants that require very specific and regular changes
in water level may be rapidly eliminated when human activity affects
the hydrology of an area. Species with highly specific dietary require-
ments are also at risk. For instance, there are species of mites that feed
only on the feathers of a single bird species. If the bird species goes
extinct, so do its associated feather mite species.
• Low tolerance for disturbance. Many species are adapted to stable, pris-
tine environments where disturbance is minimal, such as old stands
of tropical rain forests or the interiors of rich temperate deciduous
forests. When these forests are logged, grazed, burned, or otherwise
altered, these species are unable to tolerate the changed microclimatic
conditions (more light, less moisture, greater temperature variation)
or an influx of invasive species.
• Permanent or temporary aggregations. Species that group together in spe-
cific places are highly vulnerable to local extinction (Reed 1999) (see
Figure 5.12). Herds of bison, flocks of passenger pigeons, and schools
of spawning ocean fish all represent aggregations that have been ex-
ploited and overharvested by people. Many species of social animals
may be unable to persist when population size or density falls below
a certain number because they may be unable to forage, find mates,
or defend themselves.
• No prior contact with people. Species that have experienced prior human
disturbance and persisted have a lower current extinction risk than
species encountering people—along with the nonnative species as-
sociated with them—for the first time (Balmford 1996).
• Close relatives that are recently extinct or threatened with extinction. Some
groups of species, such as primates, cranes, sea turtles, and orchids,
are particularly vulnerable to extinction. The characteristics that make
certain species in these groups vulnerable are often shared by related
species.
These characteristics of extinction-prone species are not independent;
rather, they tend to group together. For example, many orchid species
have specialized habitat requirements, have specialized relationships with
pollinators, and are overharvested by collectors; all of those characteristics
lead to small, declining populations and eventually to extinction. A high
percentage of seabirds are also in danger of extinction because they have
low reproductive rates; they form dense breeding aggregations, often in
small areas, where their eggs and nestlings are prone to attack by intro-
duced predators; they are killed by oil pollution and as bycatch during
commercial fishing operations; and their eggs are overharvested by people
(Munilla et al. 2007).
By using these features to identify extinction-prone species, conserva-
tion biologists can anticipate the need for managing their populations.
Those species that are most vulnerable to extinction may have the full
range of characteristics, like the animal David Ehrenfeld (1970) imagined:
172 Chapter 5

… a large predator with a narrow habitat tolerance, long gestation


period, and few young per litter [that is] hunted for a natural product
and/or for sport, but is not subject to efficient game management. It has
a restricted distribution but travels across international boundaries.
It is intolerant of man, reproduces in aggregates, and has nonadaptive
behavioral idiosyncrasies.
There is another great gap in our knowledge of threatened species:
most threatened species that have been identified so far are members of the
best-studied groups, such as birds and mammals, highlighting the point
that only when we are knowledgeable about a species can we recognize
the dangers it faces. Other groups of species, such as beetles, ocean fish,
and fungi, are much less well known. A lack of knowledge about these
groups should not be taken to mean that the species are not threatened
with extinction; rather, it should be seen as an argument for studying those
species. The conservation status of amphibians, for example, was relatively
unknown until 15 years ago, when intensive study revealed that a high
proportion of species were in danger of extinction.

Problems of Small Populations


As mentioned above, species with small populations are in increased dan-
ger of going extinct. Small populations are subject to rapid decline in num-
bers and local extinction for three main reasons:

1. Loss of genetic diversity and related problems of inbreeding depres-


sion and genetic drift
2. Demographic fluctuations due to random variations in birth and death
rates
3. Environmental fluctuations due to variation in predation, competition,
disease, and food supply as well as natural catastrophes that occur at
irregular intervals, such as fires, floods, storms, or droughts

We’ll now examine in detail each of these causes for decline in small
populations.

Loss of genetic diversity


As described in Chapter 2, a population’s ability to adapt to a changing
environment depends on genetic variation, which occurs as a result of dif-
ferent individuals having different alleles of the same gene. Individuals
with certain alleles or combinations of alleles may have just the character-
istics needed to survive and reproduce under new conditions (Allendorf
and Luikart 2007; Frankham 2005). Within a population, the frequency of
a given allele can range from common to very rare. New alleles arise in
a population either by random mutations or through the migration of in-
dividuals from other populations. Small populations often have very low
genetic diversity, which compromises their ability to respond to environ-
Extinction Is Forever 173

mental changes. Species of yew trees (Taxus spp.) in the central


Himalayan region between Nepal and China have been reduced Because of genetic
to small populations that have such low genetic variation as to drift, small populations
be considered unsustainable (Poudel et al. 2014). lose genetic variation
In small populations, allele frequencies may change sig- more rapidly than large
nificantly from one generation to the next simply because of populations. Some small
chance—based on which individuals happen to survive to populations may lack any
sexual maturity, mate, and leave offspring. This random pro-
genetic variation.
cess of allele frequency change, known as genetic drift, is a
separate process from changes in allele frequency caused by
natural selection (Hedrick 2005). When an allele occurs at a
low frequency in the gene pool of a small population, it has
a significant probability of being lost in each generation (Figure 5.13).
For example, if a rare allele occurs in 5% of all the gene copies present in
a population of 1000 individuals, then 100 copies of the allele are present
(1000 individuals × 2 gene copies per individual × 0.05 allele frequency),
and the allele will probably remain in the population for many generations.
However, in a population of 10 individuals, only 1 copy of the allele is pres-
ent (10 individuals × 2 gene copies per individual × 0.05 allele frequency),
and it is possible that the allele will be lost from the population by chance
in the next generation.
A problem related to the loss of alleles is the reduction of heterozygos-
ity—that is, the proportion of individuals with two different alleles of a

Ne = 1000
100 Ne = 100

90
Amount of genetic variation remaining (%)

Ne = 50

80

70
Ne = 10

60

50
Ne = 5 FIGURE 5.13 The rate at which genetic varia-
40 tion is lost through genetic drift varies with popula-
tion size. This graph shows the average percentage
30
of genetic variation remaining after 10 generations
Ne = 3 in theoretical populations of various effective popu-
20
lation sizes (Ne). After 10 generations, a population
Ne = 2 with a size of 10 loses approximately 40% of its
10
genetic variation, a population of 5 loses 65%, and a
Ne = 1
population of 2 loses 95%. Blue lines indicate large
0
1 2 3 4 5 6 7 8 9 10 populations; red lines indicate small populations.
Generation (After Groom et al 2006.)
174 Chapter 5

particular gene. Considering the general case of an isolated population in


which there are two alleles of each gene in the gene pool, Wright (1931)
proposed a formula to express the proportion of the original heterozygosity
(H0; which in this case is 1) remaining after each generation (H1). The for-
mula includes the effective population size (Ne)—the size of the population
as estimated by the number of its breeding individuals2:
H1 = [1 – 1/(2Ne)] H0
According to this equation, a population of 50 breeding individuals would
retain 99% of its original heterozygosity after 1 generation:
H1 = (1 – 1/100) 1 = 1.00 – 0.01 = 0.99
The proportion of the original heterozygosity remaining after t genera-
tions (Ht/H0) decreases over time, and the decrease is greater for smaller
populations3:
Ht/H0 = (H1)t

For our population of 50 individuals, then, the remaining heterozygosity


would be 98% after 2 generations (0.99 × 0.99), 97% after 3 generations (0.99
× 0.99 × 0.99), and 90% after 10 generations (0.9910). However, a population
of 10 individuals would retain only 95% of its original heterozygosity after
1 generation, 90% after 2 generations, 86% after 3 generations, and 60%
after 10 generations. Loss of heterozygosity has been directly linked to
extinction risk (Figure 5.14).
This formula demonstrates that significant losses of genetic variation
can occur in isolated small populations. Such small populations are often
found on islands and in fragmented landscapes (Vranckx et al. 2012).
However, the amount of genetic variation within a small population can
increase over time through two means: regular mutation of genes and
migration of even a few individuals from other populations. Mutation
rates found in nature vary between 1 in 10,000 and 1 in 1 million per gene
per generation; mutations may therefore make up for genetic drift in large
populations and, to a lesser extent, in small populations. However, muta-
tions alone are not sufficient to counter genetic drift in populations of 100
individuals or fewer. Fortunately, even a low frequency of movement of
individuals between populations minimizes the loss of genetic variation
associated with small population size (Weiser et al. 2013). If even 1 or
2 immigrants arrive each generation in an isolated population of about
100 individuals, the effect of genetic drift will be greatly reduced. With
4–10 immigrants arriving per generation, the effects of genetic drift are
negligible. Gene flow from neighboring populations appears to be the

2
Factors that affect Ne, the effective population size, are discussed in detail beginning on
p. 180.
3
This is a simplification of the following equation: Ht/H0 = [1 – 1/(2Ne)]t
Extinction Is Forever 175

1.0 .99
>99% .6
Probability of extinction based on ecological variables

.3
0.9 probability .02 <2%
of extinction probability
0.8 of extinction

0.7

0.6

0.5

0.4

0.3

0.2
Extinct
0.1 Extant

0.0

0 1 2 3 4 5 6 7
Average number of heterozygous loci

Figure 5.14 Whether or not a population of the Glanville fritillary butterfly


(Melitaea cinxia) went extinct was a function of both ecological variables and
the average number of heterozygous loci; the more heterozygosity in a popula-
tion, the lower its likelihood of extinction (as shown by size of the circle). Popula-
tions that had low heterozygosity and were under high environmental pressure
(higher numbers on the y-axis) had a 100% probability of extinction, but popula-
tions subjected to high environmental pressure but which also had high het-
erozygosity had a low probability of extinction (as shown by the arrow). (After
Saccheri et al. 1998; photograph © Andrew Darrington/Alamy Stock Photo.)

major factor preventing the loss of genetic diversity in small populations


of Galápagos finches (Grant and Grant 2008). In addition, genetic variation
that increases fitness will tend to be retained longer in a population, even
when there is genetic drift.
Field data confirm that a lower effective population size leads to a
more rapid loss of alleles from a population (Evans and Sheldon 2008). For
example, in a broad survey of 89 bird species, there was a strong tendency
for abundant birds to have more heterozygosity than species with small
populations. Almost all species with populations over 10 million have over
60% heterozygosity, in contrast to most species with fewer than 100,000
individuals, which have less than 60% heterozygosity.
Unfortunately, rare and endangered species often have small, isolated
populations and thus suffer rapid losses of genetic diversity. In 170 paired

Introduction to Conservation Biology 1E Primack/Sher


Sinauer Associates
Morales Studio/SA
Primack_Sher1E_05.14 Date 1-11-16 02-17-16 2-22-16
176 Chapter 5

comparisons, threatened taxa with narrow ranges had an average of 35%


lower genetic diversity than taxonomically related nonthreatened species
with wide distributions (Spielman et al. 2004). In some cases, entire species
lacked genetic variation. In the evolutionarily isolated Wollemi pine (Wol-
lemia nobilis) of Australia, only 40 plants occur in two nearby populations.
As might be predicted, an extensive investigation failed to find any genetic
variation in this species (Peakall et al. 2003).

Consequences of reduced genetic diversity


Small populations subjected to genetic drift are susceptible to a number
of deleterious genetic effects, such as inbreeding depression, outbreeding
depression, and loss of evolutionary flexibility. These effects may con-
tribute to a decline in population size, leading to an even greater loss of
genetic diversity, a loss of fitness, and a greater probability of extinction
(Frankham et al. 2009).

inbreeding depression Matings between parents and their off-


spring, siblings, or cousins and self-fertilization in hermaphroditic spe-
cies may result in inbreeding depression, a condition that occurs when
an individual receives two identical copies of a defective allele from each
of its parents. Inbreeding depression is characterized by higher mortality
of offspring, fewer offspring, or offspring that are weak or sterile or have
low mating success (Frankham et al. 2009) (Figure 5.15). These factors
result in even fewer individuals in the next generation, leading to more
pronounced inbreeding depression.
Evidence for the existence of inbreeding depression comes from
studies of human populations (in which there are records of marriages
between close relatives for many generations), captive and wild animal
populations, and cultivated plants (Frankham et al. 2014). In a wide
range of captive mammal populations, matings among close relatives
resulted, on average, in offspring with a 33% higher mortality rate than
in non-inbred animals. This lower fitness resulting from inbreeding is
sometimes referred to as a “cost of inbreeding.” Inbreeding depression
can be a severe problem in small captive populations in zoos and in do-
mestic livestock breeding programs. Deleterious effects of inbreeding in
the wild have also been demonstrated (Crnokrak and Roff 1999): of over
150 studies, 90% showed inbreeding to be detrimental. The scarlet gilia
(Ipomopsis aggregata) provides an example: plants that come from popula-
tions with fewer than 100 individuals produce smaller seeds with a lower
rate of germination and exhibit greater susceptibility to environmental
stress than do plants from larger populations (Heschel and Paige 1995).
In another example, genetic analysis of lions from two protected areas in
Nigeria revealed evidence of inbreeding and a lack of gene flow between
the populations; in this case, trapping and moving lions as a conserva-
tion approach may be necessary to avoid the expected deleterious effects
(Tende et al. 2014).
Extinction Is Forever 177

60

50
Japanese crested ibis
Golden-cheeked warbler
50
Hatching failure (%)

40

30

20

10

0
10 100 1000 10,000 100,000
Population bottleneck size

FIGURE 5.15 Among 51 species of birds, the rate of hatching failure is high-
est for those with the smallest population sizes (as expressed by the bottleneck
population size—the lowest size recorded for the population). Hatching failure
is a function of inbreeding depression. The x-axis is on a log scale. The Japanese
crested ibis (Nipponia nippon) (top photograph) represents one extreme, with
fewer than 10 individuals in one year and hatching failure of about 45%; at the
other extreme is the golden-cheeked warbler (Dendroica chrysoparia) (bottom
photograph), with a population size always over 10,000 individuals and hatching
failure of about 5%. (After Heber and Briskie 2010; Japanese crested ibis photo-
graph © Jed Weingarten/Getty Images; golden-cheeked warbler photograph
© Rolf Nussbaumer Photography/Alamy.)

outbreeding depression Individuals of different species rarely


mate in the wild; there are strong ecological, behavioral, physiological,
and morphological isolating mechanisms that ensure that mating occurs
only between individuals of the same species. However, when a species
is rare or its habitat is damaged, outbreeding—mating between individu-
als of different populations or species—may occur. Individuals unable to
find mates of their own species may mate with individuals of related spe-
cies. The resulting offspring sometimes exhibit outbreeding depression, a
condition that results in weakness, sterility, or lack of adaptability to the
environment (Waller Biology
Introduction to Conservation 2015). 1E
Outbreeding
Primack/Sher depression may be caused by in-
compatibility
Sinauer Associatesof the chromosomes and enzyme systems that are inherited
Morales Studio/SA
from the different parents. To use an example from artificial selection, do-
Primack_Sher1E_05.15 Date 1-12-16 02-18-16 2-22-16
mestic horses and donkeys are commonly bred to produce mules. Although
178 Chapter 5

mules are not physically weak (on the contrary, they are quite strong, which
is why humans find them useful), they are almost always sterile.
Outbreeding depression can also result from matings between different
subspecies, or even matings between divergent populations of the same
species. Such matings might occur in a captive breeding program or when
individuals from different populations are kept together in captivity. In
such cases, the offspring of parents with such different genotypes are
unlikely to have the precise mixture of genes that allows individuals to
survive and reproduce successfully in a particular set of local conditions
(Frankham et al. 2009). For example, when the ibex (Capra ibex) population
of Slovakia went extinct, ibex from Austria, Turkey, and the Sinai were
brought in to start a new population. These different subspecies mated
and produced hybrids that bore their young in the harsh conditions of
winter rather than in the spring, and consequently their offspring had a
low survival rate (Templeton 1986). Outbreeding depression caused by the
pairing of individuals from the extremes of the species’ geographic range
meant failure for the experiment. However, many other studies of animals
have failed to demonstrate outbreeding depression, or have even shown
that some hybrids are more vigorous than their parent species (McClelland
and Naish 2007), a condition known as hybrid vigor. Thus, outbreeding
depression may be less of a concern for animals than inbreeding depres-
sion, the negative effects of which are well documented.
Outbreeding depression may be considerably more significant in plants,
in which the arrival of pollen on the receptive stigma of a flower is to some
degree a matter of the chance movement of pollen by wind or animal
vectors. A rare plant species growing near a closely related common spe-
cies may be overwhelmed by the pollen of the common species and fail
to produce seeds (Willi et al. 2007). Even when hybrids are produced by
matings between a common and a rare species, the genetic identity of the
rare species is lost as its small gene pool is mixed into the much larger gene
pool of the common species. The seriousness of this threat is illustrated
by the fact that more than 90% of California’s threatened and endangered
plants occur in close proximity to other species in the same genus with
which the rare plants could possibly hybridize. Such losses of identity
can also take place in gardens when individuals from different parts of a
species’ range are grown next to one another.

loss of evolutionary flexibility Genetic diversity is extremely


important to a species’ long-term survival. Rare alleles and unusual com-
binations of alleles that are harmless (or even slightly harmful) but confer
no immediate advantage on the few individuals who carry them may turn
out to be uniquely suited for a future set of environmental conditions. If
such alleles and combinations do become advantageous, their frequency
in the population will increase rapidly through natural selection because
the individuals who carry them will be those most likely to survive and re-
produce successfully, passing on the formerly rare alleles to their offspring.
Extinction Is Forever 179

Loss of genetic diversity in a small population may limit its ability to


respond to new conditions and long-term changes in the environment,
such as pollution, new diseases, or global climate change (Willi et al. 2007).
According to the fundamental theorem of natural selection, the rate of evolu-
tionary change in a population is directly related to the amount of genetic
variation in that population. A small population is less likely than a large
population to possess the genetic variation necessary for adaptation to
environmental changes and is therefore more likely to go extinct. In many
plant populations, for example, a few individuals have alleles that promote
tolerance for high concentrations of toxic metals such as zinc and lead, even
when these metals are not present. If toxic metals become abundant in the
environment because of pollution, individuals with these alleles will be
better able to adapt to them and will grow, survive, and reproduce better
than typical individuals; consequently, the frequency of these alleles in
the population will increase dramatically. However, if the population has
become small and the genotypes for metal tolerance have been lost, the
population could go extinct.

Factors that determine effective population size


Earlier in this section, we mentioned the effective population size,
which is the size of a population as estimated by the number of The effective population
its breeding individuals. The effective population size is lower size, Ne, will be much
than the total population size because many individuals do not smaller than the total
reproduce, due to factors such as inability to find a mate, being population size, N,
too old or too young to mate, poor health, sterility, malnutrition, when there is a small
and small body size (Hare et al. 2011). Many of these factors are proportion of individuals
initiated or aggravated by habitat degradation and fragmentation. reproducing, great
Furthermore, many plant, fungus, bacteria, and protist species variation in reproductive
have seeds, spores, or other structures that remain dormant in the output, an unequal sex
soil unless stable conditions for germination appear. These indi- ratio, or wide fluctuations
viduals could be counted as members of the population though in population size.
they are obviously not part of the breeding population.
Because of these factors, the effective size of a population
(Ne) is often substantially smaller than its actual size (N). Be-
cause the rate of loss of genetic diversity depends on effective population
size, loss of genetic diversity can be quite severe even in a large population.
For example, consider a population of 1000 alligators consisting of 990
immature animals and only 10 mature breeding animals: 5 males and 5
females. The effective size of this population is 10, not 1000. In a popula-
tion of a rare oak species, there might be 20 mature trees, 500 saplings,
and 2000 seedlings, resulting in a population size of 2520, but an effective
population size of only 20.
In addition, the effective population size is often even lower than the
actual number of breeding individuals because of an unequal sex ratio,
variation in reproductive output, or large annual changes in population
size (Jamieson 2011).
180 Chapter 5

unequal sex ratio A population may consist of unequal numbers


of males and females due to chance, selective mortality, or the harvesting
of one sex by people. If, for example, a population of a goose species that
is monogamous (in which one male and one female form a long-lasting
pair bond) consists of 20 males and 6 females, then only 12 individuals—6
males and 6 females—will be mating. In this case, the effective population
size is 12, not 26. In other animal species, social systems may prevent many
individuals from mating even though they are physiologically capable of
doing so. Among elephant seals, many ungulates, and many primates, for
example, a single dominant male usually mates with a large number of
females and prevents other males from mating with them (Figure 5.16),
whereas among African wild dogs and hyenas, the dominant female in
the pack often bears all of the pups.
The effect of unequal numbers of breeding males and females on Ne
can be described by this formula:
Ne = [4(Nf Nm)]/(Nf + Nm)
where Nf and Nm are the numbers of adult breeding females and adult
breeding males, respectively, in the population. In general, as the sex ratio
of breeding individuals becomes increasingly unequal, the ratio of the ef-
fective population size to the number of breeding individuals (Ne/N) goes
down. This occurs because only a few individuals of one sex are making
a disproportionately large contribution to the genetic makeup of the next
generation. In the case of Asian elephants (Elephas maximus), for example,
males are hunted by poachers for their tusks. At the Periyar Tiger Reserve
in India in 1997, there were 1166 elephants, of which 709 were adults. Of
these adults, 704 were female and 5 were male (Ramakrishnan et al. 1998).

FIGURE 5.16 Two large male


elk compete to mate with the many
smaller females surrounding them.
The effective population size is re-
duced because only one male is pro-
viding genetic input to many females.
(© Lee Foster/Alamy Stock Photo.)
Extinction Is Forever 181

If all of the adults were breeding, this sex ratio would result in an effective
population size of only 20.
In many fish and reptile species, sex is affected by temperature. As
global climate change increases water and air temperatures in many places,
the sex ratios of these species may become skewed, lowering their effective
population sizes. In Switzerland, for example, grayling (Thymallus thym-
allus) populations that used to have highly variable sex ratios centered
around 65% males before 1990 now consistently have 80%–90% males.
The effective population size for this fish species will be far lower than
the number of individuals in the population.

variation in reproductive output In many species, the number


of offspring varies substantially among individuals. This variation is par-
ticularly pronounced in highly fecund species, such as plants and fishes
(see Hedrick 2005), in which many or even most individuals produce a
few offspring while others produce huge numbers. Unequal production
of offspring leads to a substantial reduction in Ne because a few individu-
als in the present generation will be disproportionately represented in
the gene pool of the next generation. In general, the greater the variation
in reproductive output, the more the effective population size is lowered.
For a variety of species in the wild, Frankham (1995) estimated that varia-
tion in offspring number reduces effective population size by a factor of
54%. In the many annual plant populations that consist of large numbers
of tiny plants producing one or a few seeds and a few gigantic individuals
producing thousands of seeds, Ne may be reduced even more.

population fluctuations and bottlenecks In some species,


population size varies dramatically from generation to generation. Par-
ticularly good examples of this phenomenon are butterflies, annual plants,
and amphibians. In a population with extreme size fluctuations, the effec-
tive population size is much nearer the lowest than the highest number
of individuals, and it tends to be determined by the years in which the
population has the smallest numbers.
The effective size of a fluctuating population can be calculated over
a period of t years using the number of individuals (N) breeding in any
one year:
Ne = t/(1/N1 + 1/N2 + … + 1/Nt)
Consider a butterfly population, monitored for 5 years, that has 10,
20, 100, 20, and 10 breeding individuals in those successive 5 years. In
this case,
Ne = 5/(1/10 + 1/20 + 1/100 + 1/20 + 1/10) = 5/(31/100) = 5(100/31) = 16.1
The effective population size over the course of 5 years is above the
lowest population size (10), but well below the maximum (100) and the
average (32) population size.
182 Chapter 5

As these calculations suggest, a single year of drastically reduced popu-


lation numbers will substantially lower the value of Ne. This principle
applies to a phenomenon known as a population bottleneck, which occurs
when a population is greatly reduced in size. Such a population will lose
rare alleles if no individuals possessing those alleles survive and reproduce
(Jamieson 2011). With fewer alleles present and a decline in heterozygosity,
the overall fitness of the individuals in the population may decline.
A special category of population bottleneck, known as the founder ef-
fect, occurs when a few individuals leave one population and establish
a new population. The new population often has less genetic diversity
than the larger, original population. For example, the wolf population in
Sweden was established by 5 individuals (Laikre et al. 2013). Similarly,
if a population is fragmented by human activities, each of the resulting
small subpopulations may lose genetic variation and go extinct. Such is the
fate of many fish populations fragmented by dams (Wofford et al. 2005).
Bottlenecks can also occur when captive populations are established using
relatively few individuals.
The lions (Panthera leo) of Ngorongoro Crater in Tanzania provide a
well-studied example of a population bottleneck (Munson et al. 2008).
The lion population in the crater consisted of 60–75 individuals until an
outbreak of biting flies in 1962 reduced the adult population to 9 females
and 1 male (Figure 5.17). Two years later, 7 additional males immigrated
to the crater, and no new immigration was observed for the next 49 years.
The dramatic reduction in population size, the isolation of the popula-
tion, and variation in reproductive success among individuals have appar-
ently created a population bottleneck, leading to inbreeding depression.
In comparison with the large Serengeti lion population nearby, the crater
lions showed reduced genetic diversity, high levels of sperm abnormali-
ties, reduced reproductive rates, increased cub mortality, and higher rates
of disease (Munson et al. 2008). After reaching a peak of 125 animals in
the 1980s, the population declined again. By 2003, following an outbreak
of canine distemper virus that had spread from domestic dogs kept by
people living just outside the crater area, the population had dropped to
34 lions. For the next decade the population size stayed between 50 and 60
individuals, but in 2013 a coalition of four males entered the Crater. Genetic
testing will confirm whether these lions are actually “fresh blood,” but
the fact that 16 of their 20 cubs have survived—a very high rate for this
population—is a promising sign that these individuals introduced some
new alleles; the current population size is 70 lions (updates from Packer
2016, pers. comm.).
Population bottlenecks do not always lead to greatly reduced hetero-
zygosity, however. The effects of population bottlenecks are most evident
when the breeding population remains below 10 individuals for several
generations. If a population expands rapidly in size after a temporary
bottleneck, heterozygosity may be restored even though the number of
alleles present is severely reduced. An example of this phenomenon is
Extinction Is Forever 183

140 Episodes of disease


Total population size
Adult population size
120
Episode of
100 disease

80
Number

60

40

20

1962 1967 1972 1977 1982 1987 1992 1997 2002 2007 2012
Year

FIGURE 5.17 The Ngorongoro Crater lion population consisted of about


90 individuals in 1961 before crashing in 1962. Since that time, the population
reached a peak of 125 individuals in 1983 before collapsing to 34 individuals
(fewer than 20 of which were adults). Small population size, an isolated location,
lack of immigration since 1964, and the impact of disease have contributed to
the loss of genetic variation caused by a population bottleneck. A lack of cen-
sus data for certain years is the cause of gaps in the lines. The four green bars
represent episodes of disease outbreak. (After Munson et al. 2008, with updates
from C. Packer.)

the high level of heterozygosity found in the greater one-horned rhinoc-


eros (Rhinoceros unicornis) in Chitwan National Park, Nepal, even after the
population passed through a bottleneck. Population size declined from 800
individuals to fewer than 100 individuals; fewer than 30 were breeding.
With an effective population size of 30 individuals for one generation, the
population would have lost only 1.7% of its heterozygosity. As a result of
strict protection of the species by park guards, the population recovered
to 400 individuals. The Mauritius kestrel (Falco punctatus), which repre-
sents an even more extreme case, experienced a long population decline
that resulted in only one breeding pair remaining in 1974. An intensive
conservation program has allowed the population to recover to about 300
adult birds today. A study comparing the present birds with preserved
museum specimens and kestrel species living elsewhere has found that the
Mauritius kestrel lost only about half of its genetic variation after passing
through this bottleneck (Ewing et al. 2008).

managing for genetic variation These examples demonstrate that


effective population
Introduction to size is1Eoften
Conservation Biology substantially less than the total number
Primack/Sher
Sinauer Associates
Morales Studio/SA
Primack_Sher1E_05.17 Date 1-12-16 3-4-16 3-11-16
184 Chapter 5

of individuals in a population. Particularly where a combination of factors,


such as fluctuating population size, numerous nonreproductive individu-
als, and an unequal sex ratio, are involved, the effective population size
may be far lower than the number of individuals alive in a good year.
A review of a wide range of wildlife studies revealed that effective
population size averaged only 11% of total population size; that is, a popula-
tion of 300 animals, seemingly enough to maintain the population, might
have an effective size of only 33, which would indicate that it was in serious
danger of extinction (Frankham 2005). For highly fecund species, such as
fishes, seaweeds, and many invertebrates, the effective population size
may be less than 1% of the total population size (Frankham et al. 2009).
Consequently, management aimed toward simply maintaining large popu-
lations may not prevent the loss of genetic variation unless the effective
population size is also large. In the case of captive populations of rare and
endangered species, genetic variation may be effectively maintained by
controlling breeding, perhaps by subdividing the population, periodically
removing dominant males to allow subdominant males the opportunity
to mate, and periodically transporting a few selected individuals among
subpopulations.

Other factors that affect the persistence of small


populations
Random fluctuations in Random variation, or stochasticity, in the biological or physi-
birth and death rates, cal environment can cause variation in the population size of a
disruption of social species. For example, the population of an endangered butterfly
behavior by lowered species might be affected by fluctuations in the abundance of its
population density, food plants or its predators. Weather might also strongly influ-
and environmental ence the butterfly population: in an average year, the weather
stochasticity all may be warm enough for caterpillars to feed and grow, whereas
contribute to further a cold year might cause many caterpillars to become inactive
decrease in the size of and consequently starve. Such environmental stochasticity af-
populations, often leading fects all individuals in the population and is linked to demo-
to local extinction. graphic stochasticity (or demographic variation), which is varia-
tion in birth and death rates among individuals and across years
within a given population.

Demographic stochasticity
In an ideal, stable environment, a population would increase until it
reached the carrying capacity (K) of the environment, at which point the
average birthrate (b) per individual would equal the average death rate (d)
and there would be no net change in population size. In any real popula-
tion, however, individuals do not usually produce the average number of
offspring: they may leave no offspring, somewhat fewer than the average,
or more than the average. For example, in an ideal, stable giant panda (Ai-
luropoda melanoleuca) population, each female would produce an average
of two surviving offspring in her lifetime, but field studies show that rates
Extinction Is Forever 185

of reproduction among individual females vary widely around that num-


ber. However, as long as the population size is large, the average birthrate
provides an accurate description of the population. Similarly, the average
death rate in a population can be determined only by examining large
numbers of individuals because some individuals die young and other
individuals live a relatively long time. But as long as the population size
is large, the average death rate provides an accurate and relatively stable
description of the population.
Population size may fluctuate over time because of changes in the
environment or other factors without ever approaching a stable value.
Random fluctuations upward in population size are eventually bounded
by the carrying capacity of the environment, after which the popula-
tion may fluctuate downward again. In general, once population size
drops below about 50 individuals, individual variation in birth and death
rates begins to cause the population size to fluctuate randomly up or
down (Schleuning and Matthies 2009). Variation in population size due
to random variation in reproduction and mortality rates is known as
demographic variation or demographic stochasticity. If population size
fluctuates downward in any one year because of a higher than average
number of deaths or a lower than average number of births, the result-
ing smaller population will be even more susceptible to demographic
fluctuations in subsequent years.
Consequently, once a population decreases because of habitat destruc-
tion and fragmentation, demographic variation becomes important, and
the population has a higher probability of declining more and even going
extinct due to chance alone (in a year with low reproduction and high
mortality) (Melbourne and Hastings 2008). Species with highly variable
birth and death rates, such as annual plants and short-lived insects, may
be particularly susceptible to population extinction due to demographic
stochasticity. The chance of extinction is also greater in species that have
low birthrates, such as elephants, because these species take longer to
recover from chance reductions in population size.
As a simple example, imagine a population of three hermaphroditic
individuals; each lives for 1 year, attempts to find a mate and reproduce,
and then dies. Assume that each individual has a 33% probability of produc-
ing zero offspring, one offspring, or two offspring, resulting in an average
birthrate of 1 per individual; in this case, there is theoretically a stable
population. When these individuals attempt to reproduce, however, there
is a 1-in-27 chance (0.33 × 0.33 × 0.33) that no offspring will be produced in
the next generation and the population will go extinct. Consider also that
there is a 1-in-9 chance that only one offspring will be produced in the next
generation (0.33 × 0.33 × 0.33 × 3); because this individual will not be able
to find a mate, the population will be doomed to extinction in the following
generation. There is also a 22% chance that the population will decline to
two individuals in the next generation. Thus, random variation in birthrates
can lead to demographic stochasticity and extinction in small populations.
186 Chapter 5

Similarly, random fluctuations in the death rate can lead to fluctuations in


population size that could eliminate the population altogether.
When populations drop below a critical number, deviations from an
equal sex ratio may occur, leading to a declining birthrate and a further
decrease in population size. For example, imagine a population of four
birds that includes two mating pairs of males and females, in which each
female produces an average of two surviving offspring in her lifetime.
There is a 1-in-8 chance that only male or only female birds will be pro-
duced in the next generation, in which case no eggs will be laid to produce
the following generation. There is a 50% (8-in-16) chance that there will be
either three males and one female or three females and one male in the
next generation, in which case only one pair of birds will mate, and the
population will decline. This scenario is illustrated by the now extinct
dusky seaside sparrow (Ammodramus maritimus nigrescens); the last five
individuals were males, so there was no opportunity to establish a captive
breeding program. Such demographic effects are also seen in the Spanish
imperial eagle (Aquila adalberti), in which only mature birds breed when
the population is large, but immature birds are more likely to breed when
the population is small. Such immature birds are more likely to produce
predominantly male offspring, contributing to further population decline
and increasing the probability of local extinction (Ferrer et al. 2009). For
these eagles, a management strategy involving supplemental feedings was
able to increase the population size, keep mature males from leaving the
site, and restore the sex ratio.
Many small populations are demographically unstable because social
interactions (especially mating) can be disrupted once population density
falls below a certain level. This interaction among population size, popula-
tion density, population growth rate, and behavior is sometimes referred to
as the Allee effect (Allee 1931; Bonsall et al. 2014). Herds of grazing mam-
mals and flocks of birds may be unable to find food and defend themselves
against predators when their numbers fall below a certain level. Animals
that hunt in packs, such as wild dogs and lions, may need a certain number
of individuals to hunt effectively.
Perhaps the most significant aspect of the Allee effect for small popu-
lations involves reproductive behavior: many species that live in widely
dispersed populations, such as bears, spiders, and tigers, have difficulty
finding mates once the population density drops below a certain point
(Figure 5.18). Even among plant species, as population size and density
decrease, the distance between individual plants increases. Pollinating
animals may fail to visit isolated, scattered plants, resulting in insufficient
transfer of compatible pollen and a subsequent decline in seed production.
In such cases, the birthrate will decline, population density will become
lower yet, problems such as unequal sex ratio will worsen, and birth-
rates will drop even more. Once the birthrate falls to zero, extinction is
guaranteed. Detecting and anticipating Allee effects are necessary for the
management and recovery of endangered species.
Extinction Is Forever 187

Figure 5.18 Bears, such as this


Kamchatka brown bear (Ursus arctos
beringianus) in eastern Siberia, cannot
reproduce when densities decrease
below a certain number because they
are typically solitary and may have dif-
ficulty finding each other to mate.
(© Steve Winter/Getty Images.)

Environmental stochasticity and catastrophes


Random variation in the biological and physical environment, known as
environmental stochasticity, can also cause variation in the size of a popu-
lation. For example, a population of an endangered rabbit species might
be affected by fluctuations in the population of a deer species that eats the
same types of plants, in the population of a fox species that feeds on the
rabbits, and in the populations of parasites and disease-causing organ-
isms that affect the rabbits. Variation in the physical environment might
also strongly influence the rabbit population: rainfall during an average
year might encourage plant growth and allow the rabbit population to
increase, while dry years might limit plant growth and cause rabbits to
starve. Environmental stochasticity affects all individuals in the popu-
lation, unlike demographic stochasticity, which causes variation among
individuals within the population.
Natural catastrophes that occur at unpredictable intervals, such as
droughts, storms, earthquakes, and fires, along with cyclical die-offs in
the surrounding biological community, can cause dramatic fluctuations
in population levels. Natural catastrophes can kill part of a population or
even eliminate an entire population from an area. Numerous examples
exist of die-offs in populations of large mammals; in many cases, 70%–90%
of the population dies (Young 1994). For a wide range of vertebrates, the
probability of catastrophes is about 15% per generation (Reed et al. 2003).
Even though the probability of a natural catastrophe in any one year is low,
over the course of decades and centuries, natural catastrophes—includ-
ing extended periods of unseasonable weather, excessive or insufficient
rainfall, and events such as hurricanes and earthquakes—have a high
likelihood of occurring.
Modeling efforts by Menges (1992) and others have shown that random
environmental variation is generally more important than random demo-
graphic variation in increasing the probability of extinction in popula-
188 Chapter 5

tions of small to moderate size. Environmental variation can substantially


increase the risk of extinction even in populations that showed positive
population growth under the assumption of a stable environment (Mangel
and Tier 1994). In general, introducing environmental variation into popu-
lation models, in effect making them more realistic, results in populations
with lower growth rates, lower population sizes, and higher probabilities
of extinction.
Imagine a rabbit population of 100 individuals in which the average
birthrate is 0.2 and an average of 20 rabbits are eaten each year by foxes. On
average, the population will maintain its numbers at exactly 100 individu-
als, with 20 rabbits born each year and 20 rabbits eaten each year. However,
if there are 3 successive years in which the foxes eat 40 rabbits per year,
the population size will decline to 80 rabbits, 56 rabbits, and 27 rabbits in
years 1, 2, and 3, respectively. If there are then 3 years of no fox predation,
the rabbit population will increase to 32, 38, and 46 individuals in years
4, 5, and 6, respectively. Even though the same average rate of predation
(20 rabbits per year) occurred over this 6-year period, variation in year-to-
year predation rates caused the rabbit population size to decline by more
than 50%. At a population size of 46 individuals, the rabbit population
will probably go extinct within the next 5–10 years if it is subjected to the
average rate of 20 rabbits eaten by foxes per year.
The interaction between initial population size and environmental
variation was demonstrated using the biennial herb garlic mustard (Al-
liaria petiolata), an invasive plant in the United States, as an experimental
subject (Drayton and Primack 1999). Populations of various sizes were as-
signed at random either to be left alone as controls or to be experimentally
eradicated by removal of every flowering plant in each of the 4 years of
the study; removal of all plants could be considered a natural catastrophe.
Overall, the probability of an experimental population’s going extinct over
the 4-year period was 43% for small populations (fewer than 10 individuals
initially), 9% for medium-sized populations (10–50 individuals), and 7% for
large populations (more than 50 individuals). For control populations, the
probability of going extinct for small, medium, and large populations was
11%, 0%, and 0%, respectively. Large numbers of dormant seeds in the soil
apparently allowed most experimental populations to persist even when
every flowering plant was removed in 4 successive years. However, small
populations were far more susceptible to extinction than large populations.

The extinction vortex


The smaller a population becomes, the more vulnerable it is to the com-
bined effects of low genetic diversity, demographic variation, and environ-
mental stochasticity that tend to lower reproduction, increase mortality
rates, and so reduce its size even more, driving the population to extinction
(Figure 5.19). This tendency of small populations to decline toward extinc-
tion has been likened to a vortex, a whirling mass of gas or liquid spiral-
ing inward: the closer an object gets to the center of the vortex, the faster
Extinction Is Forever 189

• Environmental stochasticity
• Catastrophic events

Less genetic More Population More Lower effective


variation inbreeding more demographic population size EXTINCTION
(less ability depression subdivided by variation (Ne)
to adapt) fragmentation

• Habitat destruction
• Environmental degradation
• Habitat fragmentation
• Overharvesting
• Effects of exotic species
• Disease
• Climate change

FIGURE 5.19 Once a population drops below a certain size, it enters an


extinction vortex in which factors that affect small populations tend to drive its
size progressively lower. This downward spiral often leads to the local extinction
of species. (After Gilpin and Soulé 1986 and Guerrant 1992.)

it moves. At the center of an extinction vortex is oblivion: the


local extinction of the species. Once caught in such a vortex, Intensive management is
it is difficult for a species to resist the pull toward extinc- often required to prevent
tion (Palomares et al. 2012). Small populations often require small populations from
a careful program of population and habitat management, as declining further in size
described in later chapters, to increase their growth rates and and going extinct.
allow them to escape from the extinction vortex.

Summary
Introduction to Conservation Biology 1E Primack/Sher
Sinauer Associates „„Many extant species are on the brink of 100 and 1000 times greater than back-
Morales Studio/SA
Primack_Sher1E_05.19
extinction, with 21% of bird species, 27%
Date 1-11-16 2/23/16
ground levels. More than 99% of mod-
of mammal species, and 36% of amphib- ern species extinctions are attributable to
ian species believed to be threatened. human activity.
„„Island species have had a higher rate „„The island biogeography model is used
of extinction than mainland species. to predict the numbers of species that
Among aquatic species, freshwater spe- will persist in new protected areas and
cies apparently have a higher extinction the numbers that will go extinct else-
rate than marine species. where due to habitat destruction and
„„Current rates of extinction are between other human activities.
190 Chapter 5

„„Those species that are most vulnerable mountaintops may have locally endemic
to extinction have particular features, species with narrow distributions.
including a narrow range, one or only a „„Small populations are vulnerable to fur-
few populations, small population size, ther declines in size and eventual extinc-
declining population size, and eco- tion due to genetic, demographic, and
nomic value to humans, which leads to environmental factors, including those
overexploitation. that are stochastic (randomly occurring).
„„Rare species are more prone to extinc- Intensive management of small popula-
tion than common ones. A species can tions may be required to prevent their
be considered rare if it has one of the fol- extinction. Effective population size
lowing characteristics: it occupies a nar- measures the number of breeding indi-
row geographical range, it occupies only viduals, which is likely smaller than the
one or a few specialized habitats, or it is total population and is affected by sex
always found in small populations. Iso- ratios, age structure, and other issues.
lated habitats such as islands, lakes, and

For Discussion
1. Why should conservation biologists, or extinction? Did it have one dominant fea-
anyone else, care if a species goes locally ture that was predictive of extinction, or
extinct if it is still found somewhere else? a combination of such features?
2. Consider one of the species that has gone 3. Find out about a species that is currently
extinct in the last two centuries. Did it endangered in the wild. How might
fall into one of the categories of extinc- this species be affected by the problems
tion vulnerability listed in this chapter? of small populations? Address genetic,
What other biological or ecological traits physiological, behavioral, and ecological
might it have had that contributed to its aspects as appropriate.

Suggested Readings
Bell, C. D., J. M. Blumenthal, A. C. Broderick, and B. J. Godley. 2010. Investi-
gating potential for depensation in marine turtles: How low can you go?
Conservation Biology 24: 226–235. High gene flow may explain why certain
marine turtles can persist at low population densities.
Bonsall, M. B., C. A. Dooley, A. Kasparson, T. Brereton, D. B. Roy, and J. A.
Thomas. 2014. Allee effects and the spatial dynamics of a locally endan-
gered butterfly, the high brown fritillary. Ecological Applications 24: 108–120.
Using a modeling approach, the authors demonstrate that Allee effects
have a major effect on most populations of this butterfly.
Fisher, D. O. and S. P. Blomberg. 2012. Inferring extinction of mammals from
sighting records, threats, and biological traits. Conservation Biology 26: 57–67.
Using past observations, a model can estimate the probability that an extinct
species is still alive and that an endangered species is actually extinct.
Frankham, R., J. D. Ballou, and D. A. Briscoe. 2009. Introduction to Conservation
Genetics, 2nd Ed. Cambridge University Press, Cambridge, UK. Excellent
introduction to the importance of genetics to conservation.
Extinction Is Forever 191

Hedrick, P. 2005. Large variance in reproductive success and the Ne/N ratio.
Evolution 59: 1596–1599. A variety of factors can result in reduced effective
population size.
Morton, T. A. L., A. Thorn, J. M. Reed, R. Van Driesche, R. A. Casagrande, and
F. S. Chew. 2015. Modeling the decline and potential recovery of a native
butterfly following serial invasions by exotic species. Biological Invasion
17:1683–1695. A stochastic model is used to predict rates of loss of hetero-
zygosity in a desirable trait that allows it to adapt to an exotic host.
Palomares, F., J. A. Godoy, J. V. López-Bao, A. Rodriguez, S. Roques, M. Casas-
Marce and 2 others. 2012. Possible extinction vortex for a population of
Iberian lynx on the verge of extirpation. Conservation Biology 26: 689–697.
The Iberian lynx exhibits many of the characteristics that drive small
populations toward extinction.
Pe’er, G., M. A. Tsianou, K. W. Franz, Y. G. Matsinos, A. D. Mazaris, D. Storch,
and 5 others. 2014. Toward better application of minimum area require-
ments in conservation planning. Biological Conservation 170: 92–102. Mini-
mum area requirements have practical application in efforts to protect
populations.
Polishchuk, L. V., K. Y. Popadin, M. A. Baranova, and A. S. Kondrashov. 2015.
A genetic component of extinction risk in mammals. Oikos 124(8): 983–993.
Wake, D. B. and V. T. Vredenburg. 2008. Are we in the midst of the sixth mass
extinction? A view from the world of amphibians. Proceedings of the Na-
tional Academy of Sciences USA 105: 11466–11473. As a result of intensive
study, amphibians are now recognized as among the most threatened
animal groups.
Waller, D. M. 2015. Genetic rescue: a safe or risky bet? Molecular Ecology 24:
2595–2597. doi: 10.1111/mec.13220. Human-facilitated outcrossing be-
tween populations to increase genetic diversity can sometimes lead to
outbreeding depression.
Wedekind, C. G. Evanno, T. Székely, M. Pompini, O. Darbellay, and J. Guthruf.
2013. Persistent unequal sex ratio in a population of grayling (Salmonidae)
and possible role of temperature increase. Conservation Biology 27: 229–234.
Case study of the linkage between temperature and sex ratio in fish.
Conserving
6 Populations and
Species
Applied Population Biology 194 Prioritization: What Should Be
Conservation Categories 212 Protected? 216
Legal Protection of Species 222

This Colorado hookless cactus (Sclerocactus glaucus) is listed as


threatened under the Endangered Species Act; researchers tag
individuals within transects to track population demographics
over time.
A
t any point in time, a population of any spe-
cies can naturally be stable, increasing, de-
creasing, or fluctuating in number. In general,
widespread human disturbance destabilizes popula-
tions of many native species, often sending them into
sharp decline. But how can this disturbance be mea-
sured, and what actions should be taken to prevent or
reverse it? How can conservation biologists determine
whether a specific plan to manage a rare or endan-
gered species has a good chance of succeeding? This
chapter discusses approaches for monitoring, man-
aging, prioritizing, and protecting species and their
populations in their natural environments.
We also need to know which species to protect.
Conservation biologists and park managers do not
have enough time and money to protect every species.
Similarly, not all species need to be protected; many
species have numerous large populations that are
stable in size and cover a large area. Efforts need to be
directed to identifying species in need of protection,
using the most recent quantitative approaches that in-
clude both field data and modeling. The chapter ends
by describing how this information is used to establish
legal protection for endangered species.
194 Chapter 6

Applied Population Biology


In order to effectively protect and manage a rare or endangered species,
it is vital to have a firm grasp of its ecology, its distinctive characteristics
(sometimes called its natural history), and the status of its populations,
particularly the dynamic processes that affect population size and distri-
bution (its population biology). With more information about a rare spe-
cies, land managers can more effectively maintain it, identify factors that
place it at risk of extinction, and develop alternative management options.
Scientists who study the natural history and population biology of a
species have historically pursued this knowledge for its own sake, but
today these fields have very real and important applications in population-
level conservation efforts for rare and endangered species (see the chapter
opening photo). In many cases, however, management decisions may have
to be made without this information or while it is still being gathered.
Several types of natural history and population biology information are
important to conservation biology:
• Environment. What are the habitat types where the species is found,
and how much area is there of each? How variable is the environment
in time and space? How frequently is the environment affected by
disturbances, and of what magnitude? How have human activities
affected the environment?
• Distribution. Where is the species found in its habitat? Are individuals
clustered together, distributed at random, or spaced out regularly?
Do individuals of this species move and migrate among habitats or
to different geographical areas over short or long time periods? How
efficient is the species at colonizing new habitats? How have human
activities affected the distribution of the species or its ability to move
among habitats?
• Biotic interactions. What types of food and other resources does the spe-
cies need, and how does it obtain them? What other species compete
with it for these resources? What predators, parasites, or diseases af-
fect its population size? With what mutualists (pollinators, dispersers,
etc.) does it interact? Do juvenile stages disperse by themselves, or are
they dispersed by other species? How have human activities altered
the relationships among species in the ecosystem?
• Morphology. What does the species look like? What are the shape, size,
color, surface texture, and function of its parts? How do the shapes of
its body parts relate to their functions and help the species to survive
in its environment? What are the characteristics that allow this spe-
cies to be distinguished from species that are similar in appearance?
• Physiology. What amount of food, water, minerals, and other necessi-
ties does an individual need to survive, grow, and reproduce? How
efficient is an individual at using its resources? How vulnerable is the
species to extremes of climate, such as heat, cold, wind, and precipi-
Conserving Populations and Species 195

tation? When does the species reproduce, and what are its
special requirements during reproduction? Knowledge of the natural
• Demography. What is the current population size, and what history and population
was it in the past? Is the number of individuals stable, in- biology of a species is
creasing, or decreasing? Does the population have a mix- crucial to its protection,
ture of adults and juveniles, indicating that new individuals but urgent management
are being recruited? At what age do individuals begin to decisions often must
reproduce? be made before all this
• Behavior. How do the actions of an individual allow it to sur- information is available
vive in its environment? How do individuals in a population or while it is still being
mate and produce offspring? Do individuals of a species gathered.
interact cooperatively or competitively? How do individuals
find food? At what time of day or year is the species most
visible for monitoring?
• Genetics. How much variation occurs in morphological, physiological,
and behavioral characteristics? How is the variation spread across the
species range? How much of this variation is genetically controlled?
What percentage of the genes is variable? How many alleles does the
population have for each variable gene? Are there genetic adaptations
to local sites? Is there gene flow between populations?
• Interactions with humans. How do human activities affect the species?
What human activities are harmful or beneficial to the species? Do
people harvest or use this species in any way? What do local people
know about this species?

Methods for studying populations


Methods for studying populations have developed largely from the study
of land plants and animals (Figure 6.1). Small organisms such as protists,
bacteria, and fungi have not been investigated in comparable detail, and
species that inhabit soil, freshwater, and marine habitats are particularly
poorly investigated for population characteristics. In this section we will ex-
amine how conservation biologists undertake their studies of populations,
while recognizing that the methods need to be modified for each species.

published literature When gathering information, it is important to


remember that other people may have already investigated an ecosystem or
studied the same (or a related) species. Library indexes such as BIOSIS Cita-
tion Index and Biological Abstracts are accessible online and provide easy
access to a variety of books, articles, and reports relating to a particular
topic. This literature may contain records of previous population sizes and
distributions that can be compared with the current status of the species.
Sections of the library have material on similar topics shelved together,
so finding one book often leads to finding others. The Internet provides
ever-increasing access to databases, websites, electronic bulletin boards,
journals, news articles, specialized discussion groups, and subscription
196 Chapter 6

(A) (B)

(C)

Figure 6.1 Monitoring populations requires specialized techniques suited


to each species: (A) in rain forest research, individual trees are tagged, mapped,
and measured for their girth at 5-year intervals at painted points. (B) Ornitholo-
gists catch birds using mist-nets; record their weight, size, sex, and breeding
condition; and then attach a numbered metal band to a leg. (C) Plant ecologists
monitor changes in plant communities by placing quadrats at fixed points in a
grassland and recording the species present, their abundance, and plant height.
(A,B photographs by Richard B. Primack; C, © Paul Glendell/Alamy.)

databases such as ScienceDirect and the Thomson Reuters (formerly ISI)


Web of Science. Google Scholar is one of the best places to start searching
on topics relating to conservation biology. (When using material online,
always be sure to validate the accuracy and source of the data, especially
on sites such as Wikipedia, where there is no control over what is posted.)
Asking biologists and naturalists for ideas on references is another way to
locate published materials. Searching indexes of newspapers, magazines,
and popular journals is also an effective strategy because results of im-
portant scientific research are often covered in the popular news media.

unpublished literature Unpublished reports by scientists, enthu-


siastic citizens, government fisheries and wildlife agencies, national and
regional forest and park departments, and conservation organizations con-
Conserving Populations and Species 197

tain an enormous amount of information on conservation biology. This


“gray literature” is sometimes cited in published literature or mentioned
by leading authorities in conversations, lectures, conference presentations,
or articles. Often a report known through word of mouth can be obtained
online or through direct contact with the author. Interviews with experts
can also be used to gather their collective knowledge about endangered
species and ecosystems (McBride et al. 2012). One example is a recent study
of four threatened tree species in Fiji (Keppel et al. 2015). The research-
ers emailed questionnaires to both experts and locals to find previously
undocumented populations, improve estimates of population sizes, and
identify threats such as fire, logging, and invasive plants.

fieldwork The natural history of a species usually must be learned


through careful observations in the field (e.g., Beane et al. 2014). Fieldwork
is necessary because only a tiny percentage of the world’s species have been
adequately studied, and the ecology of a species often changes from one
place to another. Only in the field can the conservation status of a species be
determined, as well as its relationships to the biological and physical envi-
ronment. Fieldwork for species such as polar bears, humpback whales, and
tropical trees can be time-consuming, expensive, and physically arduous,
but it is crucial for developing conservation plans for endangered species,
and it can be exhilarating and deeply satisfying as well. For example, there
is a long tradition, particularly in Britain, of dedicated amateurs conducting
excellent studies of species in their immediate surroundings with minimal
equipment or financial support. However, many of the technical methods
for investigating populations are very specialized and are best learned by
studying under the supervision of an expert (Figure 6.2).
The need for fieldwork is highlighted by extensive observations of wild
penguins. One study of king penguins (Aptenodytes patagonicus) in breeding

Figure 6.2 Researchers


and volunteers release a col-
lared Argali sheep ram (Ovis
ammon) in Gun Galuut Nature
Reserve, Mongolia. The Argali
are the largest mountain sheep
in the world and are globally
endangered. Data collected from
collared animals are used to
understand range requirements
and threats of this understud-
ied species. This project is also
an example of collaboration
between a zoo and local groups.
(Photo by Richard Reading.)
198 Chapter 6

colonies on Possession Island, Crozet Archipelago, in the southern Indian


Ocean (Viblanc et al. 2014) involved fitting male penguins with data loggers
that recorded their heart rate, temperature, and overall physical activity to
better understand sources of stress on the birds during breeding. Other
research, on Magellanic penguins (Spheniscus magellanicus), used satellite
tags and radiotelemetry to define the foraging area the birds use when
feeding their chicks (Boersma and Rebstock 2009). It had previously been
thought that the birds swim out into the ocean to
forage for food only within 30 km of their nests,
but this research showed that the birds actually
swam up to 600 km and, during key periods of
chick rearing, heavily used a fishing exclusion
South America
zone (Figure 6.3). Based on this information,
the Argentinian government agreed to extend
the number of months of fishing exclusion in
this zone, and the survival of young penguins
subsequently improved.

Argentina
Golfo San Matías

Radio
transmitter

Rawson

Breeding
colonies
Atlantic Ocean

Fishing exclusion zone


0 100 Parents incubating eggs
Parents with chicks
km

Figure 6.3 Satellite tracking of Magellanic penguins (Spheniscus magellani-


cus) fitted with radio transmitters shows that penguins incubating eggs forage
up to 600 km from their breeding colonies. When penguins are feeding chicks,
foraging takes place mainly within a seasonal fishing exclusion zone that was
established to protect spawning fish. Fieldwork provided this vital information
about the penguins’ foraging habits, which led to the fishing zone remaining
closed until the chicks left their nests. (After Boersma 2006.)
Conserving Populations and Species 199

monitoring populations To learn the status of a species, scien-


tists must survey its populations in the field and monitor them over time.
Monitoring plays an important role in conservation biology. Changes in
population size and distribution can be determined by repeatedly survey-
ing a population on a regular basis (Noon et al. 2012). Efforts can also be
targeted at particularly sensitive species, such as butterflies, using them as
indicator species of the long-term stability of ecosystems (Wikström et al.
2008). The most common types of monitoring are censuses, surveys, and
population demographic studies.
The number of monitoring studies has been increasing dramatically as
government agencies have become more concerned with protecting biodi-
versity. Some of these studies are mandated by law as part of management
efforts for endangered species, and may be conducted in partnership with
conservation organizations or university researchers. Long-term monitoring
records can help to distinguish long-term population increases or declines
(possibly caused by human disturbance) from short-term fluctuations caused
by variations in weather or unpredictable natural events (Porszt et al. 2012).
Monitoring records can also determine whether an endangered species is
showing a positive response to conservation management or is responding
negatively to factors such as the present levels of harvest or the arrival of
invasive species. Observing a long-term decline in a species often provides
motivation to take vigorous action to conserve it (Figure 6.4). Field work
sometimes brings researchers into confrontations with people who are en-
gaged in illegal activities or otherwise harming species and ecosystems.

220 Figure 6.4 The American


bald eagle received rigorous
200
conservation action as a result of
180 population monitoring. Breeding
pairs observed at James Lake, Vir-
160 ginia, are shown here over time, as
Delisted in 2007 they relate to conservation activity.
Breeding pairs (N)

140 Alarmingly low numbers of nest-


ing pairs observed across the US in
120 the 1960s and 1970s spurred both
research and policy action, includ-
100
ing banning the pesticide DDT
80 (which was harming the birds
1976 officially through biomagnification), in-
1995 reduced to
listed under
60 “threatened” creasing the penalties for hunting
1972 DDT ESA
banned, and trapping eagles, and listing for
40 Pesticide protection under the Endangered
Control Act Species Act (ESA). Recovery after
20
these actions meant that the spe-
10 cies could be delisted, although it
1964 1968 1972 1976 1980 1984 1988 1992 1996 2000 2004 2008 2012 is still protected. (Inset photograph
Survey year © janbugno/Shutterstock.)
200 Chapter 6

The geographical range and intensity of monitoring have been greatly


extended through the use of volunteers, often called “citizen scientists”
(Mueller et al. 2010). Training and educating citizens not only expands
the data available to scientists but often transforms these citizens into
advocates for conservation (Matteson et al. 2012). Four programs that rely
heavily on volunteers are the North American Amphibian Monitoring
Program, Environment Canada, Project Nestwatch, and Frogwatch USA.

census A census is a count of the number of individuals present in a


population. It is a comparatively inexpensive and straightforward method,
especially for organisms that are easy to detect or are not mobile. By re-
peating a census over successive time intervals, biologists can determine
whether a population is stable, increasing, or decreasing in number. In one
monitoring study, population censuses of the Hawaiian monk seal (Mona-
chus schauinslandi) on the beaches of several islands in the Kure Atoll of the
South Pacific documented a decline from almost 100 adults in the 1950s to
fewer than 14 in the late 1960s (Figure 6.5). On the basis of these trends,

90 Mean count
Single count
80 Highest count

70 Green
Island
60
Number of seals

1960: Coast Guard


50 station opens

40 Tern
Island
30 1979: Coast Guard
station closes
20

10

0
1960 1965 1970 1975 1980 1985 1990
Year
Figure 6.5 Censusing the populations of the Hawaiian monk seal (Mona-
chus schauinslandi) on Green Island of Kure Atoll (blue line) and on Tern Island
of French Frigate Shoals (green line) revealed that this species was in danger of
extinction. Population counts were plotted from a single count, the mean of sev-
eral counts, or the maximum of several counts. Seal populations declined when
a Coast Guard station was opened on Green Island in 1960, because of distur-
bance by people and dogs; populations increased on Tern Island after the closing
of a Coast Guard station in 1979, when there was less disturbance to seals. (After
Gerrodette and Gilmartin 1990; photograph by James D. Watt, courtesy of US
Department of the Interior.)
Conserving Populations and Species 201

the Hawaiian monk seal was declared endangered in 1976 under the US
Endangered Species Act (discussed later in this chapter; Baker and Thomp-
son 2007). Subsequent conservation efforts reversed the trend, but only for
some populations. For example, the population of the Tern Island monk seal
increased after the Coast Guard station there was closed in 1979, but since
the 1990s it has substantially declined because of high juvenile mortality.
Censuses of a biological community can be conducted to determine
what species are currently present in a locality. Censuses conducted over
a wide area can help to determine the range of a species and its areas of
local abundance. As part of the North American Breeding Bird Survey
(BBS; www.pwrc.usgs.gov/bbs), thousands of participants have been re-
cording bird abundance at thousands of locations over the past 35 years
along transects, lines often designated with measuring tape or string, along
which biological data is collected (also see chapter opening photo). This
information is used to determine the stability of populations of over 400
bird species over time (Figure 6.6). A comparison of current occurrences
with past censuses can highlight species that have been lost and changes
in species ranges. These data can also be used to determine which envi-
ronmental variables (such as temperature) are important for understanding
species data (Goetz et al. 2014).

survey A survey of a population involves using a repeatable sampling


method to estimate the abundance or density of a population or species
in a part of a community. Survey methods are used when a population is
very large or its range is extensive, making a direct measure of all individu-
als impossible. An area can be divided into sampling segments and the
number of individuals in certain segments can be counted. These counts
can then be used to estimate the actual population size. For example, a
rabbit population can be surveyed by walking transects and recording the
number of rabbits observed (Dieter and Schaible 2014). Boats and planes
can be used to census whales and dolphins along transects in the ocean
(Hammond et al. 2013). New technologies are being employed to survey
difficult-to-reach species, such as the use of drones to survey populations
by taking photographs or even collecting water samples (Ore et al. 2015).
Estimating the density of organisms that are difficult to find because
they are extremely rare or very small requires special methods. Surveys
have expanded in recent years to include DNA analysis of scat and hair
samples (Hedges et al. 2013). In some cases, specially trained dogs are used
to locate scat samples of rare animal species. Such DNA studies using scats
have revealed that population size is often larger than previous estimates
suggested using traditional survey methods of direct observation, because
some of the more elusive individuals have never been seen by observers. In
an exciting new development, DNA in soil, water, and even air samples is
being used to detect the presence of rare species and the first occurrence
of invasive species. This genetic material in bulk samples is being referred
to as environmental DNA (eDNA), and it is being heralded as a new and
202 Chapter 6

(A)

(B)

Percent change per year


Less than –1.5
> –1.5 to –0.25
> –0.25 to +0.25
> +0.25 to +1.5
> +1.5

Figure 6.6 The North American Bird Survey (BBS) is one of the most com-
prehensive biological surveys in the world, as shown by this map of survey
routes. Intensity of the shading represents the density of observer routes, each
of which includes predetermined stops to record the occurrence of each bird
species. (B) A map using data from the BBS of the period covering 1966–2013 for
the grey partridge Perdix perdix (inset), for which the entire range is included
in the survey (www.mbr-pwrc.usgs.gov/bbs). Locations where the bird popula-
tions are increasing (blue) can be discerned from those where it is decreasing
(yellow and orange) and not changing (red); this information can be valuable for
protection of this species. (A, from Sauer et al. 2013; B, after Sauer et al. 2014;
inset © MikeLane45/Getty Images.)
Conserving Populations and Species 203

important conservation tool with several applications (Thomsen


and Willerlev 2015). Demographic studies
provide data on the
demographic study Demographic studies follow known numbers, ages, sexes,
individuals of different ages and sizes in a population to deter- conditions, and locations
mine their rates of growth, reproduction, and survival (Crone of individuals within
et al. 2013). Either the whole population or a subsample can be a population. These
followed. In a complete population study, all individuals are data indicate whether
counted, aged if possible, sized, sexed, and marked for future a population is stable,
identification; their position on the site is mapped; and tissue increasing, or declining
samples may be collected for genetic analysis. Techniques vary
and are the basis for
depending on the characteristics of the species and the purpose
statistical models used
of the study. Each discipline has its own techniques for follow-
to predict the future of a
ing individuals over time: ornithologists band birds’ legs, mam-
species.
malogists often attach tags to animals’ ears, and botanists nail
aluminum tags to trees; however, there is increasing concern
that for animals, the stress of being captured and tagged can
have harmful or even fatal effects (Jewell 2013). Information
from demographic studies can be used in standard mathematical formu-
las (life-history formulas) to calculate the rate of population change and to
identify critical stages in the life cycle (McCaffery et al. 2015).
An example of the application of demographic data for conservation
can be found in the work of researchers who used a combination of ge-
netic and demographic data of a threatened Mediterranean coral species
to designate and prioritize conservation units within marine protected
areas (MPAs) (Arizmendi-Mejía et al. 2015). Demographic studies can
also provide information on the age structure of a population. A stable
population typically has an age distribution with a characteristic ratio
of juveniles, young adults, and older adults. The absence or low repre-
sentation of any age class, particularly juveniles, may indicate that the
population is declining. Conversely, the presence of a large number of
juveniles and young adults may indicate that the population is stable or
even expanding.

Population viability analysis (PVA)


Monitoring data can allow us to calculate average mortality
PVA uses mathematical
rates, average recruitment rates, the current age or size distribu-
tion of the population, and the area it occupies. This informa- and statistical methods
tion can then be used to construct a mathematical model that to predict the probability
estimates the ability of a population to persist in the future, a that a population or
process known as population viability analysis (PVA). This is species will go extinct
an extension of demographic analysis (e.g., Saether and Engen within a certain time
2015). There are many mathematical and statistical models used period. PVA is also useful
for PVA, which can be thought of as risk assessment—predict- in modeling the effects of
ing the probability that a population or a species will go extinct habitat degradation and
at some point in the future (Figure 6.7). It can also be used to management efforts.
identify vulnerable stages in the natural history of the species
204 Chapter 6

Figure 6.7 Population viabil- Minimum area of habitat necessary


ity analyses predict that it takes for population to persist for 100 years
100 ha to ensure (at 95% likeli- 100

Probability population will persist (%)


hood) the persistence of a marsh Extinct population
fritillary butterfly population Extant population
80
for 100 years. All of the extinct
populations occupied areas much
smaller than 100 ha. Four of the 60
six extant populations occupy
areas smaller than 100 ha and
are predicted to go extinct in the 40
PVA
coming decade unless their habi- algorithm
tat is increased. (After Bulman curve
20
et al. 2007; photograph © Sergey
Chushkin/shutterstock.)

0 50 100 150
Amount of habitat (ha)

and to consider the effects of habitat loss, fragmentation, and deterioration


(Beissinger et al. 2009).
Another important aspect of PVA is estimating how management ef-
forts such as changing the hunting quotas or the area of protected habitat
will affect the probability of extinction (Molano-Flores and Bell 2012). PVA
can also model the effects of augmenting a population through the release
of additional individuals caught in the wild elsewhere or raised in captiv-
ity. Here are two examples of the use of PVA:
• The Hawaiian stilt (Himantopus mexicanus knudseni) is an endangered,
endemic bird of the Hawaiian Islands. Hunting and coastal develop-
ment 70 years ago reduced the number of birds to 200, but protection
has allowed recovery to the present population size of about 1600 in-
dividuals (Reed et al. 2007, 2011). The goal of government protection
efforts is to allow the population to increase to 2000 birds. A PVA was
made of the species’ ability to have a 95% chance of persisting for the
next 100 years. The model predicted that stilt numbers would increase
until they occupied all available habitat but that they would show a
rapid decline if nesting failure and mortality rates of first-year birds
exceeded 70% or if mortality rates of adults increased above 30% per
year. Keeping mortality rates below these levels would require the
control of exotic predators and the restoration of wetland habitat. And
most important, additional wetland would need to be protected if the
goal of protecting 2000 stilts was to be achieved.

• Western lowland gorillas (Gorilla gorilla gorilla) are slow-reproducing


primates that are critically endangered (Figure 6.8). Reintroduction
Introduction to Conservation Biology 1E Primack/Sher
of captive-bred and rehabilitated wild, orphaned gorillas appeared
Sinauer Associates
Morales Studio/SA
Primack_Sher1E_06.07 Date 02-10-16 2-23-16
Conserving Populations and Species 205

Figure 6.8 Population viability


analysis can be especially useful for
long-lived and slowly reproducing
species such as lowland gorillas.
(© Martin Harvey/Getty Images.)

to be successful in terms of survival of the releases, but it was not


known whether these reintroduced populations were likely to persist
over time. Demographic studies of reintroduced populations in the
Batéké Plateau region of Congo and Gabon, supplemented with data
from captive populations, were used to develop a population model
for PVA (King et al. 2014). The analysis revealed that for this species,
annual birthrates and adult female mortality rates were parameters
that particularly affected the probability of a population going extinct
within 200 years. For example, increasing the birthrate from 0.18 to
0.20 decreased the probability of extinction from 29% to 9%. Given
the current population structure, the PVA results also suggested that
these populations had a greater than 90% likelihood of surviving for
200 years. This is important support for the reintroduction program,
but it also emphasizes the necessity of monitoring over time because
parameters such as birthrate and adult female survival may change in
response to environmental stochasticity.
A key feature of PVA models, and a reason why they are especially
used with small populations (given that small populations are particularly
vulnerable to random events, as explained in Chapter 5), is that they can
incorporate demographic and environmental stochasticity. For example, if
the birthrate of a population has been determined to be 0.5, this means that
each breeding individual has a 50% chance of producing young, which is
analogous to the results of a coin flip. We know that, simply due to random
chance, when we flip a coin 10 times, we will only occasionally see exactly
5 heads and 5 tails. There is even a chance (albeit small) that we may flip 10
heads in a row. Similarly, even when we are confident of the birthrate, we
cannot be sure how many young will actually be born in any given year,
which can have especially dramatic implications for smaller populations.
Hundreds or thousands of simulations of individual populations can be
206 Chapter 6

run using this random variation to determine the probability of population


extinction within a certain period of time, the median time to extinction,
changes in population size, and changes in area occupied.
Such statistical models must be used with caution and a large dose
of common sense (Jäkäläniemi et al. 2013). Generally, about 10 years of
monitoring data are needed to obtain a PVA with good predictive power.
The results of some models can often change dramatically with different
model assumptions and slight changes in parameters. Nevertheless, PVA
is widely used and has value in demonstrating the possible effectiveness
of alternative management strategies (Sweka and Wainwright 2014). For
this reason, attempts to utilize PVA as part of practical conservation efforts
are increasingly common in management planning.

minimum viable population (mvp) PVA is also used to calculate


the minimum number of individuals of a given species required for a
population to persist over time. Some researchers argue that PVA should
be used in the establishment of recovery criteria (RC), that is, predeter-
mined thresholds that signal that an endangered species can be removed
from protection under the Endangered Species Act (Beissinger 2015). In a
groundbreaking paper, Shaffer (1981) defined the number of individuals
necessary to ensure the long-term survival of a species as the minimum
viable population (MVP): “A minimum viable population for any given spe-
cies in any given habitat is the smallest isolated population having a 99%
chance of remaining extant for 1000 years despite the foreseeable effects of
demographic, environmental, and genetic stochasticity, and natural catas-
trophes.” In other words, the MVP is the smallest population size that can
be predicted to have a very high chance of persisting for the foreseeable
future. Shaffer emphasized the tentative nature of this definition, saying
that the survival probabilities could be set at 95%, 99%, or any other per-
centage and that the time frame might similarly be adjusted, for example,
to 100 or 500 years. The key point is that the MVP size allows a quantitative
estimate to be made of how large a population must be to ensure long-term
survival. In general, protecting a larger population increases the chance of
the population persisting for a longer period of time (Figure 6.9).
One of the best-documented studies of MVP size tracked the persistence
of 120 bighorn sheep (Ovis canadensis) populations (some of which have
been followed for 70 years) in the deserts of the southwestern United States
(Berger 1990, 1999). The striking observation is that 100% of the unmanaged
populations with fewer than 50 individuals went extinct within 50 years,
while within the same time period, virtually all of the populations with
more than 100 individuals persisted. Thus, for bighorn sheep, the minimum
population size is at least 100 individuals.
Shaffer (1981) compares MVP protection efforts to flood control. It is
not sufficient to use average annual rainfall as a guideline when planning
flood-control systems; instead, we must plan for extreme situations of high
rainfall and severe flooding, which may occur only once every 50 or 100
Conserving Populations and Species 207

Goal: 90% chance of population


50% survival
surviving 1000 years
>90% survival
Requirement: 100,000 individuals
Minimum viable population size

100,000

Goal: 90% chance of population


10,000 surviving 100 years
Requirement: 3000 individuals

1000

Goal: 50% chance of population surviving 10 years


100 Requirement: 100 individuals
10 100 1000
Persistence (years)

Figure 6.9 If the goal is persistence for a greater number of years, then
a larger minimum viable population (MVP) size is needed. A greater MVP is
needed to ensure a higher probability of persistence, as illustrated in this case by
a 50% probability of survival and a greater than 90% probability of survival. Both
axes are on log scales. The values were derived from changes in population size
and persistence of 1198 species. (After Traill et al. 2010.)

years. In the same way, when attempting to protect natural systems, we un-
derstand that certain catastrophic events, such as hurricanes, earthquakes,
forest fires, epidemics, and die-offs of food items, may occur at even greater
intervals. To plan for the long-term protection of endangered species, we
must provide for their survival, not only during average years, but also
during exceptionally harsh years. Consequently, an accurate estimate of
the MVP size for a species requires an analysis of its environment. This
can be expensive and require months, or even years, of research. Analy-
ses of over 200 species for which adequate data were available (mainly
vertebrates) indicated that most MVP values for long time periods fall in
the range of 3000–5000 individuals, with a median of 4000 (Flather et al.
2011). For species with extremely variable population sizes, such as certain
invertebrates and annual plants, protecting a population of about 10,000
individuals may be the ideal strategy.
Unfortunately, many species, particularly endangered species, have
population sizes smaller than these recommended minimums. For in-
stance, half of 23 isolated elephant populations remaining in West Africa
have fewer than 200 individuals, a number considered to be inadequate
for long-term survival of the population (Bouché et al. 2011). Likewise, the
wolf population on Isle Royale, Michigan, has been fluctuating around 20
individuals but currently has only 8 adults of breeding age and no pups
(Mlot 2013).
Introduction to Conservation Biology 1E Primack/Sher
Sinauer Associates
Morales Studio/SA
Primack_Sher1E_06.09 Date 02-10-16
208 Chapter 6

Figure 6.10 Extinction rates of bird spe-


cies on the Channel Islands, with an island
scrub jay (Aphelocoma insularis) as an example
60
of one of the species. Each dot represents the
extinction percentage of all the species in that
population size class; extinction rate decreases 50
as the size of the population increases. Popu-
lations with fewer than 10 breeding pairs had 40

Extinctions (%)
an overall 39% probability of extinction over 80
years, populations of between 10 and 100 pairs 30
averaged about 10% probability of extinction,
and populations of over 100 pairs had a very
20
low probability of extinction. (After Jones and
Diamond 1976. Photograph, © Tim Zurowski/
BIA/Getty Images.) 10

0
1 10 100 1000 10,000
Population size (no. pairs)

Field evidence from long-term studies of birds on the Channel Islands


off the California coast supports the fact that large populations are needed
to ensure population persistence; only bird populations with more than
100 breeding pairs had a greater than 90% chance of surviving for 80 years
(Figure 6.10). In spite of most evidence to the contrary, however, small
populations sometimes do prevail, and many populations of birds have
survived for 80 years with 10 or fewer breeding pairs. Of course, birds are
especially mobile and can readily recolonize areas following local extinc-
tion, whereas less mobile species lack this ability.
Once an MVP size has been established for a species, the minimum dy-
namic area (MDA)—the area of suitable habitat necessary for maintaining
the minimum viable population—can be estimated by studying the size
of the home range of individuals and colonies of endangered species (Pe’er
et al. 2014; Thiollay 1989). It has been estimated that reserves in Africa of
100–1000 km2 are needed to maintain many small mammal populations
(Figure 6.11). To preserve populations of large carnivores, such as lions,
reserves of 10,000 km2 are needed.

Metapopulations
Over time, populations of a species may become extinct on a local scale
while new populations may form nearby on other suitable sites. Often a
species that lives in an ephemeral habitat, such as a streamside herb, is bet-
ter characterized in terms of a metapopulation (a “population of popula-
tions”) that is made up of a shifting mosaic of populations linked by some
degree of migration (Nöel Introduction to Conservation
et al. 2013). In some Biology 1E Primack/Sher
species, every population
Sinauer Associates
in the metapopulation isMorales
short-lived
Studio/SAand the distribution of the species
Primack_Sher1E_06.10 Date 02-10-16 2-22-16
Conserving Populations and Species 209

100,000
Small herbivores
Large herbivores
Large carnivores
Number of individuals in population

1000 animals
1000

100

0
0 1 10 100 1000 10,000 100,000 1 million 10 million
Park area (ha)

Figure 6.11 Population studies show that large parks and protected areas
in Africa contain larger populations of each species than small parks; thus, only
the largest parks may contain long-term viable populations of many vertebrate
species. Each dot represents an animal population in a park. If the viable popula-
tion size of a species is 1000 individuals (dashed line), parks of at least 100 ha
will be needed to protect small herbivores (e.g. rabbits, squirrels), parks of more
than 10,000 ha will be needed to protect large herbivores (e.g. zebras, wilde-
beests), and parks of at least 1 million ha will be needed to protect large carni-
vores (e.g., lions, hyenas). (After Schonewald-Cox 1983.)

changes dramatically with each generation. In other species,


the metapopulation may be characterized by one or more source Populations of a species
populations (core populations) with fairly stable numbers and are often connected
several sink populations (satellite populations) that fluctuate by dispersal and
in size with arrivals of immigrants. Populations in the satellite can be considered a
areas may become extinct in unfavorable years, but the areas metapopulation. In such
may be recolonized, or rescued, by migrants from the more a system, the loss of one
permanent core population when conditions become more fa- population can negatively
vorable (Figure 6.12). Metapopulations may also involve rela- affect other populations.
tively permanent populations between which individuals oc-
casionally move.
Bighorn sheep (Ovis canadensis) also offer a well-studied
example of metapopulation dynamics. These sheep have been observed
dispersing between mountain ranges and occupying previously unpopu-

Introduction to Conservation Biology 1E Primack/Sher


Sinauer Associates
Morales Studio/SA
210 Chapter 6

Figure 6.12 Possible metapopula- (A) Three independent (B) Simple metapopulation of
tion patterns, with the size of a popula- populations three interacting populations
tion indicated by the size of the circle.
The arrows indicate the direction and
intensity of migration between popula-
tions. (After White 1996.)
(C) Metapopulation with a large (D) Metapopulation with
core population and three complex interactions
satellite populations

lated sites, while mountains that previously had sheep populations are now
unoccupied. Migration and gene flow occurs primarily between popula-
tions less than 15 km apart and is greater when the intervening countryside
is hilly rather than flat (Creech et al. 2014). Maintaining dispersal routes
between existing population areas and potentially suitable sites, includ-
ing across international borders, is important in managing this species
(Buchalski et al. 2015).
The persistence of metapopulations often depends on habitat availabil-
ity. For example, destruction of the habitat of one central, core population
might result in the extinction of numerous smaller populations that depend
on the core population for periodic colonization. Effective management of a
species often requires an understanding of these metapopulation dynamics
and a restoration of lost habitat and dispersal routes.

Long-term monitoring
To understand the reasons behind population changes, monitoring of pop-
ulations needs to be combined with monitoring of other environmental
parameters. The long-term monitoring of ecosystem processes (e.g., tem-
perature, rainfall, humidity, soil acidity, water quality, discharge rates of
streams, and soil erosion) and community characteristics (species present,
percentage of vegetative cover, amount of biomass present at each trophic
level, etc.) allows scientists to determine the health of the ecosystem and
the status of species of special concern (Papworth et al. 2009). The Long-
Term Ecological Research (LTER) program in the United States focuses on
such changes on timescales ranging from months and years to decades
and centuries (Figure 6.13).
As an example of the need for long-term monitoring, certain amphibian,
insect, and annual plant populations are highly variable from year to year,
Introduction to Conservation Biology 1E Primack/Sher
Sinauer Associates
Morales Studio/SA
Primack_Sher1E_06.12 Date 02-10-16
Conserving Populations and Species 211

Research Physical Biological


Years discipline events phenomena

105 100 Millennia Evolution of


species
Paleoecology Continental Bog succession
104 10 Millennia
and limnology glaciation Forest community
migration
103 Millennium Climate Species invasion
change Forest succession
102 Century Forest fires Cultural
CO2-induced eutrophication
climate warming Population cycles
101 Decade
LTER Sun spot cycle Prairie succession
˜ events
El Nino Annual plants
100 Year Prairie fires Seasonal migration
Lake turnover Plankton
10–1 Month Ocean upwelling succession

10–2 Day Storms Algal blooms


Most ecology Daily light cycle Daily movements
Tides
10–3 Hour

Figure 6.13 The Long-Term Ecological Research (LTER) program focuses


on timescales ranging from months to centuries in order to understand changes
in the structure, function, and processes of ecosystems that are not apparent
from short-term observations. (From Magnuson 1990.)

so many years of data are required to determine whether a particular spe-


cies is actually declining in abundance over time or merely experiencing a
number of low population years that are in accord with its regular pattern
of variation. In one instance, more than 50 years of observation of popula-
tions of two flamingo species (Phoenicopterus ruber, the greater flamingo,
and Phoeniconaias minor, the lesser flamingo) in southern Africa revealed
that large numbers of chicks fledged only in years with high rainfall, mak-
ing it appear that reproduction was simply highly variable (Figure 6.14).
In fact, there was a 31-year gap in which no major hatchings occurred and
the population appeared to be in peril. Fortunately, the population began
to increase in 2008 and 2011, when numerous chicks hatched.
The fact that environmental effects may lag for many years behind their
initial causes creates a challenge to understanding change in ecosystems.
For example, acid rain, nitrogen deposition, and other components of air
pollution may gradually change the water chemistry, algal community, and
oxygen content of forest streams, ultimately making the aquatic environ-
ment unsuitable for the larvae of certain insect species. In this case, the
cause (air pollution) may have occurred years or even decades before the
effect (insect decline) becomes detectable. Even habitat fragmentation can
have delayed effects on losses via gradual environmental degradation and
metapopulation extinction.

Introduction to Conservation Biology 1E Primack/Sher


212 Chapter 6

800

700

600
Rainfall (mm/yr)

500

400

300

200

100

0
1956 1960 1964 1968 1972 1976 1980 1984 1988 1992 1996 2000 2004 2008 2012
Year

Figure 6.14 The bars show rainfall data from Etosha Na-
tional Park in southern Africa for the years 1956–2013. The fla-
mingo breeding events that occurred in those years are indicated
by circles. Orange circles indicate failed breeding events; eggs
wither laid but not chicks hatched. The small, medium, and large
green circles indicate, respectively, fewer than 100 chicks hatched,
hundreds of chicks hatched, and thousands of chicks hatched.
There was a 31-year gap between 1976 and 2008 in which no large
hatching even occurred. (After Simmons 1996, with updates from
R. E. Simmons. Photograph © Kevin Schafer/DigitalVision/Photoli-
brary.com.)

A major purpose of monitoring programs is to gather essential data


on biological communities and ecosystem functions that can be used to
document changes over time. Monitoring in these studies allows manag-
ers to determine whether the goals of their projects are being achieved
or whether adjustments must be made in the management plans (called
adaptive management), as discussed in Chapter 10.

Conservation Categories
The IUCN uses quantit- Once the data collected on populations and species allows us to
ative information, includ- identify those species most vulnerable to extinction, it is useful
ing the area occupied to create a system whereby extinction risk can be categorized to
by the species and facilitate the prioritization of conservation efforts. To mark the
the number of mature status of rare and endangered species for conservation purpos-
individuals presently es, the International Union for Conservation of Nature (IUCN)
alive, totoassign
Introduction species
Conservation Biologyto has established conservation categories (Figure 6.15) (www.
1E Primack/Sher
Sinauer Associates categories.
conservation iucn.org). These categories have proved useful in establishing
Morales Studio/SA
Primack_Sher1E_06.14 protection
Date 02-10-16 2-24-16 for threatened species at the national and interna-
Conserving Populations and Species 213

Extinct
Extinct (EX)
Extinct in the wild (EW)

Threatened
Critically endangered (CR)
Adequate data Endangered (EN)
Vulnerable (VU)

Evaluated

Lower
Near threatened (NT)

risk
Least concern (LC)
Species Inadequate data
Data deficient (DD)

Not evaluated (NE)

Figure 6.15 The IUCN categories of conservation status. This chart shows
the distribution of the categories. Reading from left to right, they depend on
(1) whether a species has been evaluated or not and (2) how much information
is available for the species. If data are available, the species is then put into a
category of lower risk, threatened, or extinct. (After IUCN 2001.)

tional levels and directing attention toward species of special concern. The
conservation categories follow:
• Extinct (EX). The species (or other taxon, such as subspecies or variety)
is no longer known to exist. The IUCN currently lists 709 animal spe-
cies and 90 plants species as extinct.
• Extinct in the wild (EW). The species exists only in cultivation, in captiv-
ity, or as a naturalized population well outside its original range. The
IUCN currently lists 32 animal species and 29 plant species as extinct
in the wild.
• Critically endangered (CR). The species has an extremely high risk of
going extinct in the wild, according to any of the criteria A–E (Table
6.1).
• Endangered (EN). The species has a very high risk of extinction in the
wild, according to any of the criteria A–E.
• Vulnerable (VU). The species has a high risk of extinction in the wild,
according to any of the criteria A–E.
• Near threatened (NT). The species is close to qualifying for a threatened
category but is not currently considered threatened.
• Least concern (LC). The species is not considered near threatened or
threatened. (Widespread and abundant species are included in this
category.)
Introduction to Conservation
• Data deficient (DD).Biology 1E Primack/Sher
Inadequate information exists to determine the risk
Sinauer Associates
of extinction
Morales Studio/SA for the species.
Primack_Sher1E_06.15 Date 02-10-16
214 Chapter 6

TABLE 6.1 IUCN Red List Criteria for the Assignment of Conservation Categories
Quantification of criteria for Red List category “critically
Red List criteria A–E endangered”a
A. Observable reduction in The population has declined by 80% or more over the last 10 years
numbers of individuals or three generations (whichever is longer), either based on direct
observation or inferred from factors such as levels of exploitation,
threats from introduced species and disease, or habitat destruction
or degradation.
B. Total geographical area occupied The species has a restricted range (<100 km2 at a single location)
by the species and there is observed or predicted habitat loss, fragmentation,
ecological imbalance, or heavy commercial exploitation.
C. Predicted decline in number of The total population size is less than 250 mature, breeding individuals
individuals and is expected to decline by 25% or more within 3 years or 1
generation.
D. Number of mature individuals The population size is less than 50 mature individuals.
currently alive
E. Probability the species will go Extinction probability is greater than 50% within 10 years or 3
extinct within a certain number generations.
of years or generations
a
A species that meets the described quantities for any one of criteria A–E may be classified as critically endangered. Simi-
lar quantification for the Red List categories “endangered” and “vulnerable” can be found at www.iucnredlist.org.

• Not evaluated (NE). The species has not yet been evaluated against the
Red List criteria.
When used on a national or other regional level, there are two additional
Red List categories:
• Regionally extinct (RE). The species no longer exists within the country
(region) but is extant in other parts of the world.
• Not applicable (NA). The species is not eligible for the regional Red List
because, for example, it is not within its natural range in the region (it
has been introduced) or because it is only a rare migrant to the region.
Species in the critically endangered, endangered, and vulnerable cat-
egories are considered threatened with extinction. For these three catego-
ries, the IUCN has developed quantitative measures of threat based on the
probability of extinction. These Red List criteria, described in
Table 6.1, are based on the developing methods of PVA. These
criteria focus on population trends and habitat condition. The
The IUCN system has advantage of this system is that it provides a standard method
been used to identify of classification by which decisions can be reviewed and evalu-
Red Lists of threatened ated according to accepted quantitative criteria, using whatever
species and to deter- information is available.
mine whether species Using habitat loss as a criterion in assigning categories is
are responding to particularly useful for many species that are poorly known
conservation efforts. biologically, because species can be listed as threatened if their
habitat is being destroyed even if scientists know little else about
Conserving Populations and Species 215

them. In practice, a species is most commonly assigned to an IUCN cat-


egory based on the area it occupies, the number of mature individuals it
has, or the rate of decline of the habitat or population; the probability of
extinction is least commonly used (van Swaay et al. 2011).
Using the criteria in Table 6.1 and the categories in Figure 6.15, the
IUCN has evaluated and described the threats to plant and animal species
in its series of Red Data Books and Red Lists of threatened species; these
detailed lists of endangered species by group and by country can be seen at
www.iucn.org. Species listed as threatened include 1462 of 5499 described
mammal species, 2097 of 10,052 bird species, and 2303 of 6338 amphibian
species (see Table 5.1).
Most bird, amphibian, and mammal species have been evaluated using
the IUCN system, but the levels of evaluation are lower for reptiles, fish,
and flowering plants, resulting in low apparent levels of threat. Even
though numerous species of fish (2523), reptiles (1160), molluscs (2197),
insects (1259), crustaceans (1735), and higher plants (10,686) are designated
as threatened with extinction, most species in these groups have still not
yet been evaluated. The evaluations of insects and other invertebrates,
mosses, algae, fungi, and microorganisms are even less adequate (Régnier
et al. 2009). While in most cases the lack of data leads to underestimates
of extinction risk, it has been argued that for some groups the risk may
be overestimated when based on only presence–absence data, such as for
amphibians (Cruickshank et al. 2016).
By tracking the conservation status of species over time, it is possible
to determine whether species are responding to conservation efforts or are
continuing to be threatened (Lacher et al. 2012). One such measure is the
Red List Index, which demonstrates that the conservation status of certain
animal groups has continued to decline since 1988, with particularly sharp
declines for albatrosses, petrels, and many amphibians (Baillie et al. 2008;
Quayle et al. 2007). Within Australia, the Red List Index is declining faster
than the Global Red List Index, indicating the need for additional conserva-
tion action (Szabo et al. 2012). Another measure, the Living Planet Index,
follows population sizes for 2688 vertebrate species; this index has shown
an average decline of 61% for tropical species and an average 31% increase
for temperate species from 1970 to 2008.
A program similar to the efforts of the IUCN is the NatureServe net-
work of Natural Heritage programs, which covers all 50 US states, 3 Cana-
dian provinces, and 14 Latin American countries (www.natureserve.org/
explorer). This network, strongly supported by The Nature Conservancy,
gathers, organizes, and manages information on the occurrence of “ele-
ments of conservation interest”—more than 64,000 species, subspecies,
and biological communities, in addition to half a million precisely located
populations (Figure 6.16) (De Grammont and Cuarón 2006). Elements
are given status ranks based on a series of standard criteria: number of
remaining populations or occurrences, number of individuals remaining
(for species) or extent of area (for communities), number of protected sites,
216 Chapter 6

70
Presumed/possibly extinct
60
Critically imperiled

50 Imperiled
Percentage of species
Vulnerable
40

30

20

10

0
69% 51% 45% 42% 36% 33% 30% 23% 21% 20% 19% 18% 18% 15%

ies

als

s
ls

rs

ts

ies

tle

ile

er

rd
ian
fis

fis
se

lan
pe

all

lfl

m
pp

er

Bi
pt
ee
ay
us

ib

er
op

se

am
p

sp
e
rb

ki
rn
ph
m

Cr

at

R
g

am
sh

no
/s

M
fe
in
hw

ge
er

Am
as

s/

m
ies

/d
er
at

Ti
Gr

es

Gy
ow

rn
hw

rfl

es
Fr

Fe

fli
tte
Fl
es

on
Bu
Fr

ag
Dr
More imperiled Less imperiled

Figure 6.16 Some species groups from the United States ranked as presumed
extinct, critically imperiled, imperiled, or vulnerable according to criteria endorsed
by The Nature Conservancy and coordinated by NatureServe. The groups are ar-
ranged with those at greatest risk on the left. Freshwater species are at greater
risk of extinction than terrestrial species. (After Wilcove and Master 2005.)

degree of threat, and innate vulnerability of the species or community.


When methods of the Natural Heritage system and the IUCN categories
are used to evaluate the same species, the resulting rankings of threat
are quite similar. On the basis of these criteria, elements are assigned an
imperilment rank from 1 to 5, ranging from critically imperiled (1) to de-
monstrably secure (5), on a global, national, and regional basis. Species
are also classified as “X” (extinct), “H” (known historically with searches
ongoing), and “unknown” (uninvestigated). Data on these conservation
elements are available on the NatureServe website (www.natureserve.org).

Prioritization: What Should Be Protected?


Although some conservationists would argue that no species should ever
be lost, the
Introduction to Conservation reality
Biology is that numerous species are in danger of going extinct
1E Primack/Sher
Sinauer Associates
and there are too few resources available to save them all. The challenge
Morales Studio/SA
to conservation
Primack_Sher1E_06.16 Dateefforts
2-8-16 lies in finding ways to minimize the loss of biodi-
versity during a period of limited financial and human resources (Watson
et al. 2011).
Conserving Populations and Species 217

the species approach One approach to establishing conservation


priorities involves protecting particular species—and in doing so, protect-
ing an entire biological community and associated ecosystem processes
(Branton and Richardson 2011). Protected areas are often established to pro-
tect individual species of special concern, such as rare species, endangered
species, keystone species, and culturally significant species. Species such
as these, which provide the impetus to protect an area and ecosystems, are
known as focal species. One type of focal species is an indicator species,
a species that is associated with an endangered biological community or
set of unique ecosystem processes; for instance, the endangered northern
spotted owl is a forest indicator species in the Pacific Northwest of the
United States. Many national parks have also been created to protect flag-
ship species, such as tigers and pandas, which capture public attention,
have symbolic value, and are crucial to ecotourism. Flagship and indica-
tor species are also known as umbrella species because protecting them
automatically protects other species and aspects of biodiversity.
Another way in which the prioritization of species for conservation can
be determined is by evaluating the degree to which it is DUE (distinctive,
utilitarian, and endangered) (Figure 6.17):
1. Distinctiveness (or irreplaceability). A species is often given high con-
servation value if it is taxonomically distinctive—that is, it is the only
species in its genus or family (Faith 2008). Similarly, a population of a
species having unusual genetic characteristics that distinguish it from
other populations of the species might be a high priority for conserva-
tion. An ecosystem composed primarily of rare endemic species or with
other unusual attributes (small area, scenic value, unique geological
features) is given a high priority for conservation.
2. Utility. Species that have present or potential value to people, such as
wild relatives of wheat, are given high conservation priority. Species
with major cultural significance, such as tigers in India and the bald
eagle in the United States, are given high priority.

Figure 6.17 The Komodo dragon


(Varanus komodoensis) of Indonesia
is an example of a species that fits
all three categories: it is the world’s
largest lizard (distinctiveness), it has
major potential as a tourist attraction
in addition to being of great scientific
interest (utility), and it occurs on only a
few small islands of a rapidly develop-
ing nation (endangerment). (© Barry
Kusuma/Getty Images.)
218 Chapter 6

3. Endangerment (or vulnerability). Species in danger of extinction are of


greater concern than species that are not; thus, the whooping crane
(Grus americana), with only about 382 individuals, requires more pro-
tection than the sandhill crane (Grus canadensis), with approximately
520,000 individuals.
The species approach follows from creating survival plans for individual
species, which are developed by governments and private conservation or-
ganizations. In the Americas, the Natural Heritage Programs and the Na-
tureServe network use information on rare and endangered species to target
new localities for conservation—areas where there are concentrations of
endangered species or where the last populations of a declining species exist
(www.natureserve.org). In the IUCN Species Survival Commission, over 100
specialist groups provide action plans for endangered animals and plants.

the ecosystem approach A number of conservationists have ar-


gued that rather than species, ecosystems and the biological communi-
ties they contain should be targeted for conservation (Tallis and Polasky
2009). They claim that spending $1 million or so on habitat protection and
management of a self-maintaining ecosystem might preserve more spe-
cies and provide more value to people in the long run than spending the
same amount of money on an intensive effort to save just one conspicu-
ous species (Figure 6.18). It often is easy to demonstrate an ecosystem’s
economic value to policy makers and the public in terms of flood control,
clean water, and recreation, whereas arguing for a particular species may
be more difficult. Thus, combining species and ecosystem approaches may
be a good conservation strategy in many circumstances.
When using this ecosystem-based approach, conservation planners
should try to ensure that representative sites of as many types of ecosys-
tems as possible are protected. A representative site includes the species
and environmental conditions characteristic of the ecosystem. Although
no site is perfectly representative, biologists working in the field can often
identify suitable sites for protection.
Where immediate decisions must be made to determine park bound-
aries and which species and ecosystems to protect, biologists are being
trained to make rapid biodiversity assessments, also known as RAPs (for
rapid assessment programs). RAPs involve mapping vegetation, making
lists of species, checking for species of special concern, estimating the total
number of species, and looking out for new species and features of special
interest. A bioblitz (explained in Chapter 2) is one type of RAP.

the wilderness approach Wilderness areas are a related priority,


in part because they are more likely to contain species and populations
that need protecting, such as old-growth trees (Figure 6.19). Large blocks
of land that have been minimally affected by human activity, have a low
human population density, and are not likely to be developed in the near
Conserving Populations and Species 219

(A) Figure 6.18 (A) The Monterey


Bay National Marine Sanctuary
in California was established to
protect a coastal marine environ-
ment that includes marine mam-
mals, seabirds, and ocean bottom
species. (B) Feather worms and
various coral species living on the
ocean bottom at the sanctuary.
(Photographs courtesy of NOAA.)

(B)

future are also perhaps the only places in the world where large mammals
can survive in the wild. These wilderness areas can also serve as reference
areas for restoration (see Chapter 10). It is worth emphasizing that even
these so-called wilderness areas have had a long history of human activity
and people have often affected the structure of the biological communi-
ties they contain.

the hotspot approach Certain organisms can be used as biodiver-


sity indicators to highlight new areas where concentrations of species can
be protected. For example, a site with a high diversity of flowering plants
often, but not always, will also have a high diversity of mosses, spiders,
and fungi. Further, areas with high diversity often have a high percent-
220 Chapter 6

Figure 6.19 Old-growth giant sequoias are protected in this designated


wilderness area within Sequoia and Kings Canyon National Park. (© Neale Clark/
robertharding/Getty Images.)

age of endemism—species occurring there and nowhere else (Joppa et al.


2011). The IUCN Plant Conservation office in England has documented
about 250 global centers of plant diversity with large concentrations of
species (Hoffmann et al. 2008). In a similar effort, BirdLife International
has identified over 200 Important Bird Areas (IBAs) with more than 2400
restricted-range bird species; many of these localities are in urgent need
of protection (www.birdlife.org).
Using a similar approach, Conservation International, World Wildlife
Fund, and others have designated hotspots that have great biological di-
versity and high levels of endemism and that are under immediate threat
of species extinctions and habitat destruction (www.biodiversitya-z.org/
content/biodiversity-hotspots) (Figure 6.20). Using these criteria, 34 global
hotspots have been targeted for new protected areas. These hotspots to-
gether encompass the entire ranges of 12,066 endemic species of terrestrial
vertebrates (42% of the world’s total) and at least part of the ranges of an
additional 35% of the remaining terrestrial vertebrate species—all on only
2.3% of Earth’s total land surface.
One major center of biodiversity is the tropical Andes, where 30,000
plant species, 1728 bird species, 569 mammal species, 610 reptile species,
and 1155 amphibian species persist in tropical forests and high-altitude
grasslands on about 0.3% of Earth’s total land surface. The hotspot ap-
proach has generated a considerable amount of enthusiasm and funding,
and it will be worth watching to see how successful it is in advancing the
goals of conservation in areas of intense human pressures, some of which
are sites of armed conflicts (Hanson et al. 2009).
Conserving Populations and Species 221

(A)

Western Ghats
and Sri Lanka Indo-Burma
Caribbean
Philippines
Polynesia Wallacea
and East
Micronesia Melanesia
S Guinean C N
Mesoamerica forests
Western Colombia Sundaland
and Ecuador
Atlantic forests
Madagascar and the New
Tropical Andes of Brazil
Indian Ocean islands Caledonia
East African
coastal forests

(B)

Mountains of
Caucasus central Asia
California
region Himalaya Japan
Mediterranean
Basin
Iran- Mountains of
Mexican Anatolia southwest China
woodlands Afromontane

Brazilian Horn of
Cerrado Africa
South
African Eastern
Karoo South Africa
Chilean
temperate Cape Floral region Southwest
forest Australia
New Zealand

Figure 6.20 Hotspots are targets for protection because of their high
biodiversity, endemism, and significant threat of imminent extinctions. (A) Six-
teen tropical rain forest hotspots. Areas circled in green are island groups. The
Polynesia/Micronesia region (far left) covers a large number of Pacific Ocean
islands, including the Hawaiian Islands, Fiji, Samoa, French Polynesia, and the
Marianas. Black-circled letters indicate the only three remaining undisturbed
rain forest areas of any extent, in South America (S), the Congo basin of Africa
(C), and the island of New Guinea (N). (B) Eighteen hotspots representing other
ecosystems. Yellow dots denote areas that have experienced armed conflicts
between 1950 and 2000 with over 1000 casualties. (After Hansen et al. 2009;
Mittermeier et al. 2005.)
222 Chapter 6

Legal Protection of Species


Once conservation biologists have identified a species as need-
National governments ing protection, laws can be passed and treaties can be signed to
protect designated implement conservation efforts. National laws protect species
endangered species within individual countries, while international agreements pro-
within their borders, vide a broader framework for conservation.
establish national parks,
and enforce legislation on
National laws
environmental protection. People in many countries recognize that preserving a healthy
environment and protecting species are linked to sustaining
human health. National governments and national conservation
organizations in such countries acknowledge this and play an
important role in the protection of all levels of biological diversity. Laws
are passed to establish national parks and other protected areas; to regulate
activities such as fishing, logging, and grazing; and to limit air and water
pollution. International treaties that restrict trade in endangered animals
are implemented at the national level and enforced at the borders. The
true measure of a nation’s commitment to protecting biodiversity is the
effectiveness with which these laws are enforced.
In European countries, endangered species conservation is accom-
plished through domestic enforcement of international agreements such
as the Convention on International Trade in Endangered Species (CITES)
and the Ramsar Convention on Wetlands (see the section “Intenational
agreements” on p. 226). Species that occur on the IUCN’s international
Red Lists of endangered species and in national Red Data books are also
protected (Fontaine et al. 2007). To indicate where these species can be
found, the Fauna Europaea database provides information on the distri-
bution of 130,000 terrestrial and freshwater species. Countries in Europe
protect species and habitats through directives adopted by the European
Union; these directives implement the earlier Bern Convention, which was
established to protect endangered species in Europe, with a special focus
on migratory species. Some countries may have additional laws, such as
the Wildlife and Countryside Act of 1981 in the United Kingdom, which
protects habitat occupied by endangered species.
Despite the fact that many countries have enacted legislation to preserve
biodiversity, it also is true that national governments are sometimes unre-
sponsive to requests from conservation groups to protect the environment.
In some cases national governments have acted to decentralize decision
making, relinquishing control of natural resources and protected areas
to local governments, village councils, and conservation organizations.

the us endangered species act The United States has developed


a system for designating and protecting threatened species that is separate
from the IUCN system. In the United States, the principal conservation law
protecting species is the Endangered Species Act (ESA), passed in 1973 and
subsequently amended in 1978 and 1982. The ESA was created by the US
Congress to “provide a means whereby the ecosystems upon which en-
Conserving Populations and Species 223

dangered species and threatened species depend may be conserved [and]


to provide a program for the conservation of such species.” Species are
protected under the ESA if they are on the official list of endangered and
threatened species. In addition, a recovery plan is generally required for
each listed species (Himes Boor 2014).
As defined by law, “endangered species” are those likely to become
extinct as a result of human activities or natural causes in all or a sig-
nificant portion of their range; “threatened species” are those likely to
become endangered in the near future. The secretary of the Interior De-
partment, acting through the US Fish and Wildlife Service (FWS), and
the secretary of the Commerce Department, acting through the National
Marine Fisheries Service (NMFS), can add and remove species from the
list based on information available to them. Since 1973, more than 1519
species in the United States have been added to the list, including many
well-known species such as the whooping crane (Grus americana) and
the manatee (Trichechus manatus), in addition to 625 endangered species
from elsewhere in the world that face special restrictions when they are
imported into the United States.
Many species are listed under the ESA only when they have fewer than
100 individuals remaining, making recovery difficult (see Chapter 5). An
early listing of a declining species might allow it to recover and thus become
a candidate for removal from the list sooner than if authorities wait for its
status to worsen before adding it to the list. The great majority of
species in the United States listed under the ESA are flowering
plants (839 species) and vertebrates (403 species), despite the fact The ESA mandates such
that most of the world’s species are insects and other invertebrates strong protection for
(ecos.fws.gov/tess_public/reports/box-score-report). Only 70 in- species that conservation
sect species are currently listed; however, if the same proportion and business groups
of insects as of vertebrates were protected, an estimated 29,000 often agree to compro-
species would be protected under the ESA, an awesome number
mises that allow some
to contemplate (Dunn 2005). Clearly, greater efforts must be made
species protection
to study the lesser known and underappreciated invertebrate
along with limited
groups and extend listing to those endangered species whenever
development.
necessary (Stankey and Shindler 2006).
The protection afforded to species listed under the ESA is so
strong that business interests and landowners often lobby strenu-
ously against listing species in their area. At the extreme are landowners
who destroy endangered species or species being considered for listing
on their property to evade the provisions of the ESA, a practice informally
known as “shoot, shovel, and shut up.” Such was the fate of the threat-
ened Preble’s meadow jumping mouse (Zapus hudsonius), which lives in
streamside habitats in Colorado and Wyoming on a quarter of the sites that
contained suitable habitat (Brook et al. 2003). Many argue that landowners
should be compensated in some way to provide an important incentive to
conserve endangered species and their habitats. Financial incentives exist
to help private landowners improve habitats for endangered species (see
Chapter 11), but they may be too small to be effective in many cases.
224 Chapter 6

Another important obstacle to listing is the difficulty of species recov-


ery—rehabilitating species or reducing the threats to species to the point
where they can be removed from listing under the ESA, or “delisted.” So far,
only about 30 of more than 1436 listed species in the United States have been
delisted because of recovery, and another 25 species have shown enough
recovery to be reclassified from endangered to threatened (Schwartz 2008).
The most notable successes include the brown pelican (Pelecanus occidentalis),
the American peregrine falcon (Falco peregrinus), and the American alligator
(Alligator mississippiensis). In 2007, the bald eagle (Haliaeetus leucocephalus)
was removed from the federal list of threatened and endangered species
because its numbers in the lower 48 states had increased from 400 breeding
pairs in the 1960s to over 9000 pairs (see Figure 6.4).
Overall, most listed species are still declining in range and abundance,
and unfortunately, for around 20% of species there is insufficient data to
determine whether their populations are changing over time (Leidner and
Neel 2011). Due to their low numbers and consequent vulnerability, there
is now recognition that even species that are candidates for delisting will
still require some degree of conservation management to maintain their
populations (Redford et al. 2011) (Figure 6.21).
The difficulty of implementing recovery plans for so many species is
often not primarily biological but, rather, political, administrative, and

Independent

Adapted to human-dominated Peregrine falcon


environments Gray wolf

Maintains viable populations under American alligator


existing regulatory mechanisms Gray whale
Brown pelican
Stages of recovery

Periodic intervention Kirtland’s warbler


Continuous intervention to eliminate Robbins’ cinquefoil
or decrease a limiting factor Hawaiian gallinule
Continuous intervention to restore
desirable ecological processes at Salmon and other
landscape level large river fishes

Sustained in wild as a result California condor


of captive releases
Only occurs in captivity Guam kingfisher
Loulu palm

Dependent

Figure 6.21 Endangered species often require active management and


intervention as part of the recovery process. There is a continuum, with some
species independent of humans and others dependent on human intervention.
(After Scott et al. 2005.)
Conserving Populations and Species 225

ultimately financial (Briggs 2009). For example, an endangered river clam


species might need to be protected from pollution and the effects of an
existing dam. Installing sewage treatment facilities and removing a dam
are theoretically straightforward actions but they are expensive and dif-
ficult to carry out in practice. Total expenditures reported for fiscal year
2013 were $1.75 billion, of which $1.67 billion was reported by federal agen-
cies and $76 million was reported by the states (USFWS 2013). However,
this amount of funding is less than the amount requested for listing and
recovery purposes (Gibbs and Currie 2012). The cost would be much higher
if the US government granted private landowners financial compensation
for ESA-imposed restrictions on the use of their property, an option that
is periodically discussed in the US Congress.
Funding for the ESA has been growing steadily over the past 20 years,
but the number of protected species has been growing even faster. As a
result, there is less money available per species. The importance of ad-
equate funding for species recovery is shown by a study demonstrating
that species that receive a higher proportion of requested funding for their
recovery plans have a higher probability of reaching a stable or improved
status than species that receive a lower proportion of funding (Miller
et al. 2002). The longer a species has been protected under the ESA, the
higher is the probability that it is improving in status (Taylor et al. 2005)
(Figure 6.22). Also, species have a higher probability of improving if
critical habitat and a recovery plan have been designated for them (Gibbs
and Currie 2012).
Concerns about the implications of ESA protection force business orga-
nizations, conservation groups, and goverments to develop compromises

80

70

60
Proportion of species (%)

Continuing to decline
50
Figure 6.22 The longer spe-
cies have been listed, protected,
and managed under the Endan-
40
gered Species Act, the greater is
their probability of improving in
30 status (as shown by the whoop-
Improving ing crane) and the slower is their
20 probability of continuing to de-
cline in status (with the Indiana
10 bat as an example). The numbers
do not add up to 100% because
0 some species are not changing in
0 5 10 15 20 25 30 status and others are of unknown
Time listed (years) status. (After Taylor et al. 2005.)
226 Chapter 6

that reconcile both conservation and business interests (Camacho 2007). To


provide a legal mechanism to achieve this goal, Congress amended the ESA
in 1982 to allow the design of habitat conservation plans (HCPs). HCPs are
regional plans that allow development in designated areas but also protect
remnants of biological communities or ecosystems that contain groups of
actual or potentially endangered species. These plans are drawn up by the
concerned parties—developers, conservation groups, citizen groups, and
local governments—and given final approval by the FWS. About 650 HCPs
covering about 16 million ha and over 500 species have been approved. In
one case, an innovative program in Riverside County, California, allows
developers to build within the historic range of the endangered Stephens’
kangaroo rat (Dipodomys stephensi) if they contribute to a fund that will be
used to buy land and establish wildlife sanctuaries to protect the species.
Already, more than $42 million has been used to secure 41,000 ha, with
the long-term goal of raising $100 million. As a result of the HCP and the
resulting new reserves, the Stephens’ kangaroo rat is being considered for
removal from the endangered species list; however, the long-term effective-
ness of these measures remains to be seen. While HCPs are not perfect,
they are attempts to create the next generation of conservation planning.
They seek to protect many species, entire ecosystems, or whole communi-
ties; extend over a wide geographical region; and include many projects,
landowners, and jurisdictions.

international agreements International agreements to protect


species and other aspects of biodiversity and ecosystems are needed for
several reasons: species migrate across borders, there is international trade
in biological products, the benefits of biodiversity are of international im-
portance, and the threats to biodiversity are often international in scope.
International agreements have provided a framework for countries to co-
operate in protecting species, ecosystems, and genetic variation. Treaties
are negotiated at international conferences and come into force when they
are ratified by a certain number of countries, often under the authority of
international bodies, such as the United Nations Environment Programme
(UNEP), the Food and Agriculture Organization of the United Nations
(FAO), and the IUCN.
The benefits of biodiversity conservation also accrue globally, as the
environment can supply natural products for agriculture, medicine, and
industry and genetic materials for research and breeding new or stronger
varieties of species for human use. Tropical regions are also important
in the global ecosystem through their influence on carbon dioxide (CO2)
levels and weather patterns. Each country is ultimately responsible for
protecting its own natural environment, but until their economies become
stronger, developing countries may be unable to pay for the habitat pres-
ervation, research, and management required for the task. Because bio-
diversity conservation is important at both the national and global levels,
it is fair for the developed countries of the world (including the United
Conserving Populations and Species 227

States, Canada, Japan, Australia, and many European nations) to help pay
to protect biodiversity.
The protection of biodiversity must be addressed at multiple levels
of government. Although the major control mechanisms that presently
exist in the world are based within individual countries, international
agreements among countries are increasingly used to protect species, eco-
systems, and genetic variation. International cooperation is an absolute
requirement for several reasons:
• Species migrate across international borders. Conservation efforts must
protect species at all points in their ranges; efforts in one country will
be ineffective if critical habitats are destroyed in a second country to
which an animal migrates (Bradshaw et al. 2008) (see Chapter 8). For
example, efforts to protect migratory bird species in northern Europe
will not work if the birds’ overwintering habitat in Africa is destroyed.
Efforts to protect whales in US coastal waters will not be effective
if these species are killed or harmed in international waters. Species
are particularly vulnerable when they are migrating, as they may be
more conspicuous, more tired, or more desperately in need of food and
water. Globally, international parks, often called “peace parks,” have
been created to protect species living and moving through border areas,
such as the Waterton–Glacier International Peace Park on the border
of the United States and Canada, which protects grizzly bears (Ursus
arctos horribilis) and lynx (Lynx canadensis).
• International trade in biological products is commonplace. A strong demand
for a product in one country can result in the overexploitation of the
species in another country to supply this demand. When people are
willing to pay high prices for exotic pets, plants, or wildlife products
such as rhino horn, poachers looking for easy profits, or poor, desperate
people looking for any source of income, will take or kill even the very
last animal to obtain this income. To prevent overexploitation, consum-
ers who buy wildlife products, and the people who collect and trade
them, need to be educated about the consequences of overuse of wild
species. When poverty is the root of overexploitation, it is sometimes
possible to provide people with economic alternatives while strictly
controlling resource use (see Chapter 11). Where exploitation stems
from greedy people seeking to make a profit, laws and enforcement
efforts such as border checks should be strengthened.
• Biodiversity provides internationally important benefits. The community of
nations benefits from the species and genetic variation used in agricul-
ture, medicine, and industry; the ecosystems that help regulate climate;
and the national parks and other protected areas of international sci-
entific and tourist value. It is also widely recognized that biodiversity
has intrinsic value, existence value, and option value (see Chapter 3).
The developed countries of the world that use and rely on biodiversity
and ecosystem services from poor tropical countries provide limited,
228 Chapter 6

inadequate funding to help these less-wealthy countries manage and


protect globally significant resources. Funding levels need to be in-
creased and the funds must be used more effectively.
• Many environmental pollution problems that threaten ecosystems are interna-
tional in scope. Such threats include atmospheric pollution and acid rain;
the pollution of lakes, rivers, and oceans; greenhouse gas production
and global climate change; and ozone depletion (Srinivasan et al. 2008).
Additionally, the environmental costs of many of these problems do not
fall on countries in proportion to their role in causing them. For exam-
ple, the United States and China are the world’s leading producers of
greenhouses gases (see Chapter 4), but many low-lying countries such
as Bangladesh and the Maldives will be most affected by the rising sea
levels associated with climate change. Or consider the River Danube,
which carries the pollution of a vast agricultural and industrial region
that spans 10 countries before it empties into the Black Sea, another
international body of water bordered by 4 additional countries. Prob-
lems such as these can only be solved by countries working together.

International agreements to protect species


To address the protection of biodiversity, countries worldwide have signed
a number of key international agreements. These agreements have pro-
vided a framework for countries to cooperate in protecting species, habi-
tats, ecosystem processes, and genetic variation. Treaties are negotiated at
international conferences and come into force when they are ratified by
a certain number of countries, often under the authority of international
bodies such as the UNEP, the FAO, and the IUCN.
As was mentioned earlier, one of the most important treaties protect-
ing species at an international level is the Convention on International
Trade in Endangered Species (CITES), established in 1973 in association
with UNEP (Reeve 2014; www.cites.org). Currently there are 179 member
countries. CITES, headquartered in Switzerland, establishes lists (known
as Appendices) of species for which international trade is to be controlled
or monitored. Member countries agree to restrict trade in these
species and halt their destructive exploitation. Regulated plants
include important horticultural species such as orchids, cycads,
CITES has developed
cacti, carnivorous plants, and tree ferns; timber species and
extensive lists of
wild-collected seeds are increasingly being considered for regu-
species for which lation as well. Closely regulated animal groups include parrots,
trade is prohibited, large cats, whales, sea turtles, birds of prey, rhinos, bears, and
controlled, or monitored. primates. Species collected for the pet, zoo, and aquarium trades
Many countries have and species harvested for their fur, skin, or other commercial
established their own products are also closely monitored.
Red Data list of species International treaties such as CITES are implemented when a
that they protect within country signing the treaty passes laws to enforce it. Nongovern-
their borders. mental organizations (NGOs) such as the IUCN, Wildlife Trade
Program, the TRAFFIC network run by the World Wildlife Fund
Conserving Populations and Species 229

(WWF) and the IUCN, and UNEP’s World Conservation Monitoring Centre
(WCMC) provide technical advice regarding legal and enforcement aspects
of CITES to national governments. Countries may also protect species listed
by national Red Data books. Once species protection laws are passed within
a country, police, customs inspectors, wildlife officers, and other govern-
ment agents can arrest and prosecute individuals possessing or trading in
protected species and can seize the products or organisms involved (Figure
6.23). For example, in November 2013, Thai authorities at the Bangkok airport
seized hundreds of endangered turtles being shipped illegally in passenger
baggage. What made this story particularly unusual is that the seizure and
news story took place just a day after a major CITES conference in Bangkok.
Member countries are required to establish their own management
and scientific authorities to implement CITES obligations within their
own borders (see Chapter 11). CITES is particularly active in encouraging
cooperation among countries, in addition to fostering conservation efforts
by development agencies. The CITES Secretariat periodically sends out bul-
letins aimed at publicizing specific illegal activities. For example, in recent
years, the CITES Secretariat has recommended that its member countries

Figure 6.23 A fur reference collection in northern China. For some prod-
ucts, such as the zebra skin, the type of animal involved can be easy to identify,
but for other products, such as bags, coats, rugs, and shoes, the type of animal
used to make them may be hard to determine, and often requires microscopic
analysis of hairs. (Photograph by Richard Primack.)
230 Chapter 6

Figure 6.24 The burning of ivory in Kenya. To keep ivory off the internation-
al market and hopefully reduce the killing of wild elephants, wildlife authorities
in Kenya burned more than 15 tons of elephant tusks seized from poachers in
2015, and made plans to burn an addtional 120 tons of both elephant tusks and
rhino horns in 2016. (© Carl de Souza/AFP/Getty Images.)

halt wildlife trade with Vietnam because of that country’s unwillingness


to restrict the illegal export of wildlife from its territory.
CITES has been instrumental in restricting trade in certain endangered
wildlife species. Its most notable success was a global ban on the ivory
trade after poaching caused severe declines in African elephant popula-
tions (Figure 6.24). Recently, countries in southern Africa with increasing
elephant populations have been allowed to resume limited ivory sales,
resulting in an unfortunate increase in illegal harvesting.
While illegal trade in wildlife may not sound important, it is a huge
business, estimated to be between $10 and $20 billion per year, excluding
aquatic species (see Chapter 4). It remains a major problem and is sometimes
linked to illegal trade in timber, drug smuggling, and arms trafficking. Not
surprisingly, compiling accurate data on illegal wildlife trade is a challenge
(Sajeva et al. 2013). One difficulty with enforcing CITES is that shipments of
both living and preserved plants and animals are often mislabeled, due to
either ignorance of species names or deliberate attempts to avoid the restric-
tions of the treaty. Also, sometimes countries fail to enforce the restrictions
of the treaty because of corruption or a lack of trained staff. Finally, many
restrictions are difficult to enforce because some international borders are
remote, rugged, and difficult to monitor, such as that between Laos and
Vietnam. As a result, the illegal wildlife trade continues to pose one of the
most serious threats to biodiversity, particularly in Asia.
Conserving Populations and Species 231

Another key treaty is the Convention on the Conservation of Migratory


Species of Wild Animals (CMS), often referred to as the Bonn Convention,
which focuses primarily on bird species (www.cms.int). This conven-
tion, which has been signed by 119 countries, complements CITES by
encouraging international efforts to conserve bird species that migrate
across international borders and by emphasizing regional approaches
to research, management, and hunting regulations. The convention now
includes protection of bats and their habitats and cetaceans in the Baltic
and North Seas. Other important international agreements that protect
species include:
• Convention on the Conservation of Antarctic Marine Living Resources
(www.ccamlr.org)
• International Convention for the Regulation of Whaling, which estab-
lished the International Whaling Commission (www.iwc.int)
• International Convention for the Protection of Birds and the Benelux
(Belgium/Netherlands/Luxembourg) Convention Concerning Hunt-
ing and the Protection of Birds
• Convention for the Conservation and Management of Highly Mi-
gratory Fish Stocks in the Western and Central Pacific Ocean (www.
wcpfc.int)
• Additional agreements protecting specific groups of animals, such as
prawns, lobsters, crabs, fur seals, Antarctic seals, salmon, and vicuña
A number of more broadly focused international agreements are also
increasingly seeking direct protection of endangered species. For example,
the Convention on Biological Diversity, described in Chapter 11, now in-
cludes recommendations for the protection of IUCN Red Listed species
(www.iucnredlist.org).
A weakness of all these international treaties is that they operate
through consensus, so strong measures often are not adopted if one or
more countries oppose them. Also, any nation’s participation is volun-
tary, meaning that countries can choose to ignore these conventions and
pursue their own interests if they find the conditions of compliance too
difficult (Carraro et al. 2006). This flaw was highlighted when several
countries decided not to comply with the International Whaling Com-
mission’s 1986 ban on whale hunting, and the Japanese government an-
nounced that it would continue hunting whales under the dubious claim
that further data were needed to evaluate the status of whale populations.
Persuasion and public pressure are the principal means used to induce
countries to enforce treaty provisions and prosecute violators, though
funding through treaty organizations can also help. An additional prob-
lem is that many conventions are underfunded and are consequently
ineffective in achieving their goals. Unfortunately, there are often no
monitoring mechanisms in place to determine whether countries are
even enforcing the treaties.
232 Chapter 6

Summary
„„Protecting and managing a rare or en- ries: extinct, extinct in the wild, critically
dangered species requires a firm grasp endangered, endangered, vulnerable,
of its ecology and its distinctive char- near threatened, least concern, data defi-
acteristics (sometimes called its natural cient, and not evaluated.
history). Long-term monitoring of a „„Priorities for protection can be deter-
species in the field can determine if it is mined in several ways, including the
stable, increasing, or declining in abun- species approach, the ecosystem ap-
dance over time. proach, the wilderness approach, and
„„Population viability analysis (PVA) the hotspot approach.
uses demographic, genetic, and envi- „„National governments protect biodiver-
ronmental data to estimate how vari- sity by establishing national parks and
ous management actions will affect the refuges, controlling imports and exports
probability that a population will persist at their borders, and creating regula-
until some future date. It can be used to tions for air and water pollution. The
calculate the minimum viable popula- most effective law in the United States
tion (MVP) size: the smallest population for protecting species is the Endangered
size that can be predicted to have a high Species Act (ESA).
chance of persisting for the foreseeable
„„At the international level, the Conven-
future. The MVP for many species is at
tion on International Trade in Endan-
least several thousand individuals.
gered Species (CITES) allows gov-
„„A species may be best described as a ernments to regulate, monitor, and
metapopulation made up of a shifting sometimes prohibit trade in individuals
mosaic of populations that are linked by and products from endangered species.
some degree of migration.
„„The IUCN has developed quantitative
criteria for populations and ecosystems
to assign species to conservation catego-

For Discussion
1. How might you monitor populations of and a hotspot approach for protecting it.
a species of fish over time? Would your Which approach will be most effective
methods differ if the fish was a freshwa- for protecting this species? Which ap-
ter species or a marine species? Why or proach is most feasible? Which approach
why not? will protect the most species in addition
2. Choose a threatened species in your re- to your target?
gion. Weigh the merits and limitations of 3. A wide range of laws protect endangered
a species-centered approach, an ecosys- species. Why don’t species covered by
tem approach, a wilderness approach, such laws quickly recover?
Conserving Populations and Species 233

Suggested Readings
Carroll, C., R. J. Frederickson, and R. C. Lacy. 2014. Developing metapopu-
lation connectivity criteria from genetic and habitat data to recover the
endangered Mexican wolf. Conservation Biology 28: 76–86. Metapopulation
models demonstrate that dispersal between populations is crucial to pre-
venting local extinction for small populations.
Douglas, L. R. and K. Alie. 2014. High-value natural resources: Linking wild-
life conservation to international conflict, insecurity, and development con-
cerns. Biological Conservation 171: 270–277. Social, economic, and political
issues must be addressed for conservation goals to be achieved.
Duarte, A., J. S. Hatfield, T. M. Swannack, M. R. Forstner, et al. 2016. Simulat-
ing range-wide population and breeding habitat dynamics for an endan-
gered woodland warbler in the face of uncertainty. Ecological Modelling 320:
52–61. Mathematical tools that incorporate stochasticity assist conservation
management planning.
Hedges, S., A. Johnson, M. Ahlering, M. Tyson, and L. S. Eggert. 2013. Accura-
cy, precision, and cost-effectiveness of conventional dung density and fecal
DNA based survey methods to estimate Asian elephant (Elephas maximus)
population size and structure. Biological Conservation 159: 101–108. New
DNA methods are greatly improving our ability to estimate population
size.
Jäkäläniemi, A., A. H. Postila, and J. Tuomi. 2013. Accuracy of short-term de-
mographic data in projecting long-term fate of populations. Conservation
Biology 27: 552–559. Many years of data are needed to build a good PVA
model.
Liu, P., L. Sun, J. Li, L. Wang, et al. 2015. Population viability analysis of Gloy-
dius shedaoensis from northeastern China: A contribution to the assessment
of the conservation and management status of an endangered species.
Asian Herpetological Research 1(1): 48–56. PVAs have been used to evaluate
current protection and management strategies for Chinese snakes.
NatureServe. 2009. http://natureserve.org. This website organizes and pres-
ents data on biodiversity surveys from North America.
Pittman, S. E., M. S. Osbourn, and R. D. Semlitsch. 2014. Movement ecology of
amphibians: A missing component for understanding population declines.
Biological Conservation 169: 44–53. Studies of juvenile and adult movement
patterns are critical for the conservation management of amphibians.
Schwartz, M. W. 2008. The performance of the Endangered Species Act. Annual
Review of Ecology, Evolution, and Systematics 39: 279–299. Many listed spe-
cies are recovering, but certain goals have not been achieved.
Thomsen, P. F. and W. Willerslev. 2015. Environmental DNA—An emerging
tool in conservation for monitoring past and present biodiversity. Biologi-
cal Conservation 183: 4–18. Detecting whole organisms or just traces of
organisms from bulk samples of soil, water, or air expands our capacity to
survey populations and species diversity.
Bringing Species
7 Back from the Brink
Establishing and Reinforcing Can Technology Bring Back
Populations 236 Extinct Species? 261
Ex Situ Conservation
Strategies 246

European bison, or
wisent (Bison bonasus),
at snowfall, Germany.
I
n Chapters 5 and 6, we discussed the problems
conservation biologists face in preserving
naturally occurring populations of endan-
gered species. This chapter discusses some exciting
conservation methods used to establish new wild and
semi-wild populations of rare and endangered spe-
cies and increase the sizes of existing populations.
These methods include breeding species in zoos,
aquaria, and botanic gardens—organizations that also
assist conservation through education and research
programs. Captive breeding and other approaches to
augment or establish new populations may allow spe-
cies that have persisted only in captivity or in small,
isolated populations to regain their ecological and
evolutionary roles within their ecosystems. Further-
more, simply increasing the number and size of its
populations generally lowers the probability a species
will go extinct.
Population establishment programs are unlikely to
be effective, however, unless the factors leading to
the decline of the original wild populations are clear-
ly understood and eliminated, or at least controlled
(Venevsky et al. 2005). For example, endangered am-
phibians are increasingly being raised in captivity,
236 Chapter 7

but they cannot be released back into the wild if they lack resis-
Establishing new popul- tance to the chytrid fungal pathogen Batrachochytrium dendro-
ations of endangered batidis, a nonnative and now widespread fungus that is killing
species can benefit the them worldwide (Kolby et al. 2015). One possibility is to breed
species itself, other captive amphibian populations for fungal resistance before at-
species, and the eco- tempting to establish new populations in the wild.
system. However, such
programs must identify Establishing and Reinforcing
and eliminate the factors Populations
that led to the original
Three basic approaches, all involving relocation of existing cap-
population’s decline.
tive-bred or wild-collected individuals, have been used to estab-
lish new populations of animals and plants and to enlarge exist-
ing populations. The IUCN’s Reintroduction Specialist Group
coordinates many of these efforts (www.iucnsscrsg.org):
• A reintroduction program1 involves releasing captive-bred or wild-col-
lected individuals at an ecologically suitable site within their historical
range where the species no longer occurs (Carter et al. 2016).
• A reinforcement program involves releasing individuals into an exist-
ing population to increase its size and gene pool (Smyser et al. 2013);
this approach can be thought of as restocking or augmentation. These
released individuals may be raised in captivity or may be wild indi-
viduals collected elsewhere.
• An introduction program involves moving captive-bred or wild-col-
lected animals or plants to areas suitable for the species outside their
historical range.

The main objectives of a reintroduction program are to create a new


population in its original environment and to help restore a damaged
ecosystem. For example, a program initiated in 1995 to reintroduce gray
wolves (Canis lupus) into Yellowstone National Park aims to restore the
equilibrium of predators, herbivores, and plants that existed prior to
human intervention in the region (Figure 7.1A). Wild-collected animals
are also sometimes caught and then released elsewhere within the range
of their species when a new protected area has been established, when
an existing population is under a new threat in its present location, or
when natural or artificial barriers to the normal dispersal tendencies of
the species exist. If possible, individuals are released near the site where
they or their ancestors were collected to ensure genetic adaptation to
their environment.

1
Some confusion exists about the terms denoting the establishment of populations. Rein-
troduction programs are sometimes called reestablishments or restorations. Another term,
translocation, usually refers to moving individuals from a location where they are about to
be destroyed to another site that, hopefully, provides a greater degree of protection.
Bringing Species Back from the Brink 237

(A)

(B)

1600
Wolf population
Wolves killed
1400 Domestic animals killed

1200
Number of individuals

1000

800

600

400

Wolf reintroduction
200

0
19
19
19
19
19
19
19
19
19
19
19
19
19
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
87
88
89
90
91
92
93
94
95
96
97
98
99
00
01
02
03
04
05
06
07
08
09
10
11
12
13
14

Year

Figure 7.1 (A) A gray wolf in Yellowstone National Park wears a radio
transmitter collar that allows researchers to follow its movements. (B) The
number of wolves in Wyoming, Idaho, and Montana increased following the
reintroduction of wolves to the Yellowstone area in 1995. There has also been an
increase in the number of domestic animals, primarily sheep, killed by wolves,
and an increase in the number of problem wolves killed by government authori-
ties. (A, photograph courtesy of William Campbell/US Fish and Wildlife Service;
B, after Musiani et al. 2003, with updates courtesy of M. Musiani, and from Clark
and Johnson 2009.)
238 Chapter 7

In contrast to reintroduction, introduction to an entirely new location


outside the existing range of a species may be appropriate when the envi-
ronment within the known range of a species has deteriorated to the point
at which the species can no longer survive there, or when reintroduction is
impossible because the factor causing the original decline is still present
in the species’ range. This approach is sometimes referred to as assisted
colonization (Lunt et al. 2013). The New Zealand kakapo, for example, was
introduced to offshore islands because nonnative predators had decimated
its populations on the mainland (Moro et al. 2015). In the near future,
introductions may be necessary for many species if those species can no
longer survive within their current ranges because of a warming climate
(Hendricks et al. 2016).
The reintroduction or introduction of a species must be carefully
thought out so that the released species does not damage its new ecosys-
tem or harm populations of any local endangered species (Olden et al.
2011). Care must be taken that released individuals have not acquired any
diseases that could spread to and decimate wild populations.
One special method used in establishing new populations or reinforc-
ing existing populations is head-starting, an approach in which animals are
raised in captivity during their vulnerable young stages and then released
into the wild. The release of sea turtle hatchlings produced from eggs col-
lected from the wild and raised in nearby hatcheries is an example of this
approach (see Chapter 1).

Considerations for animal programs


Establishing new populations is often expensive and difficult because it
requires a serious, long-term commitment. The programs that capture,
raise, release, and monitor sea turtles, peregrine falcons, and whooping
cranes, for example, have cost millions of dollars and have required years of
work. When the animals involved are long-lived, the program may have to
continue for many years before its outcome is known (Grenier et al. 2007).
Population establishment programs can also become highly emotional
public issues, as demonstrated by the programs for the California con-
dor, the black-footed ferret, and the gray wolf in the United States and
comparable programs in Europe. These programs are often criticized
on many different fronts. They may be attacked as a waste of money
(“Millions of dollars for a few ugly birds!”), unnecessary (“Why do we
need wolves here when there are so many elsewhere?”), intrusive (“We
just want to go about our lives without the government telling us what
to do!”), poorly run (“Look at all the ferrets that died of disease in captiv-
ity!”), or unethical (“Why can’t the last animals just be allowed to live
out their lives in peace without being captured and put into zoos?”). The
answer to all of these criticisms is straightforward. Although not appro-
priate for every endangered species, a well-run, well-designed captive
breeding and population establishment program may be the best hope
for a species’ preservation.
Bringing Species Back from the Brink 239

Because of the conflicts and high emotions involved, it is crucial that


population establishment programs include local people so that (ideally)
the community has a stake in the program’s success. (Indeed, this is true
of any conservation project.) At a minimum, it is necessary to explain the
need for the program and its goals and to convince local people to support
it—or at least not to oppose it (Yochim and Lowry 2016). Programs are
often more successful if they provide incentives to affected people rather
than imposing rigid restrictions and laws. For example, direct payments
are made to Wyoming residents whose domestic animals are injured by
reintroduced wolves, and the few wolves that repeatedly attack livestock
are either killed or moved in order to retain local support for the program
(Boitani et al. 2011) (Figure 7.1B).
Successful reintroduction programs often have considerable educational
value. In Brazil, conservation and reintroduction efforts for the golden lion
tamarin (Leontopithecus rosalia) have become a rallying point for the protec-
tion of the last remaining fragments of the Atlantic forest (see textbook
cover image). In the Middle East and northern Africa, captive-bred Arabian
oryx (Oryx leucoryx) have been successfully reintroduced into many desert
areas that they formerly occupied, providing a source of national pride and
opportunities for employment (Fisher 2016).
The introduction of a species to new sites needs to be carefully considered
and evaluated in order to ensure that the species does not damage its new
ecosystem or harm any local populations of other endangered species (Olden
et al. 2011). Care must be taken to ensure that released individuals have not
acquired any diseases while in captivity that could spread to and decimate
wild populations. For example, captive black-footed ferrets (Mustela nigripes)
must be carefully handled and quarantined so that they do not acquire
diseases from people or dogs that they might transfer into wild populations
upon their release in North American grasslands (Figure 7.2A).
There is an important genetic component to the selection of plants or
animals for reintroduction programs. Captive populations may have lost
much of their genetic variation or, when raised for several generations
in captive conditions, may have become genetically adapted to the be-
nign captive environment. Such genetic changes, which lower the species’
ability to survive in the wild following release, have occurred in captive
populations of Pacific salmon. Individuals have to be carefully selected to
guard against inbreeding depression and to produce a genetically diverse
release population (Ottewell et al. 2014). Wild-caught individuals must be
selected from a population living in an environment and climate that is as
similar as possible to that of the release site. After release, a species may
adapt genetically to its new environment such that the gene pool from the
original wild population is not actually being preserved.
Some animal species may require special care and assistance immedi-
ately after release to increase their survival prospects (Harrington et al.
2013; White et al. 2012). This approach is known as soft release. Animals
may have to be fed and sheltered at the release point until they are able
240 Chapter 7

to subsist on their own, or they may need to be caged temporarily at


the release point and introduced gradually, once they become familiar
with the sights, sounds, smells, and layout of the area (Figure 7.2B).
Eighty-eight chicks of the Mauritius kestrel (Falco punctatus) raised in
captivity by humans and then given a soft release into the wild on the
Indian Ocean island of Mauritius, for example, had a survival rate that
was not significantly different from that of 284 chicks born in the wild
(Nicoll et al. 2004).
Animals can also be released without assistance such as food supple-
mentation (hard release), although reintroductions of this type are more
likely to succeed with wild-caught over captive-bred individuals (Bocci

(A)

(B)

Figure 7.2 (A) A black-footed ferret


raised at the captive colony in Colorado.
(B) Cages allow black-footed ferrets to
experience the environment into which
they will eventually be released. The
ferrets’ caretaker is wearing a mask to
reduce the chance of exposing the ani-
mals to human diseases. (A, photograph
by Ryan Hagerty/US Fish and Wildlife
Service; B, photograph by M. R. Match-
ett, courtesy of US Fish and Wildlife
Service.)
Bringing Species Back from the Brink 241

et al. 2016). Intervention may be necessary if animals appear unable to


survive, particularly during episodes of drought or low food abundance
(Blanco et al. 2011). Even when animals appear to have enough food to
survive, supplemental feeding may help by increasing reproduction and
allowing the population to persist and grow. Outbreaks of diseases and
pests may have to be monitored and dealt with. The effects of human ac-
tivities in the area, such as farming and hunting, need to be observed and
possibly controlled. In every case, a decision has to be made about whether
it is better to give occasional temporary help to the species or to force the
individuals to survive on their own (Harrington et al. 2013).
Establishment programs for common wildlife species managed for
hunting have always been widespread and have contributed a great deal of
knowledge to biologists trying to establish new populations of endangered
species. A number of generalizations can be made from analyses of about
200 establishment programs for such common game species (Fischer and
Lindenmayer 2000; Hughes and Lee 2015):
• Success was greater for releases in excellent-quality habitat (84%) than
in poor-quality habitat (38%).
• Success was greater in the core of the historical range (78%) than at the
periphery of and outside the historical range (48%).
• Success was greater with wild-caught (75%) than with captive-reared
animals (38%).
• Success was greater for herbivores (77%) than for carnivores (48%).
• For the bird and mammal species studied, the probability of establish-
ing a new population increased with the number of animals being re-
leased, up to about 100 individuals. Releasing more than 100 animals
did not further enhance the probability of success.
Other examinations of the published literature have found that the suc-
cess rate of reintroduction projects involving endangered mammals, birds,
reptiles, amphibians, and fish is generally much lower than that of projects
involving wildlife managed for hunting, suggesting that improved meth-
ods need to be developed for endangered species (Griffiths and Pavajeau
2008; Harding et al. 2015). In fact, the rates of success may be even lower
than these values indicate because the results of many projects, particularly
those that fail, are not published and are poorly documented (Miller et al.
2014; White et al. 2012). As we’ll see later in the chapter, this low rate of
success emphasizes that protecting existing populations of endangered
species and improving their habitat must remain our highest priorities.
Clearly, monitoring of establishment programs is crucial in determining
whether the programs are achieving their stated goals (Seddon et al 2014;
McCleery et al. 2014). Monitoring may need to be carried out over many
years, even decades, because many reintroductions that initially appear
successful eventually fail. The key elements of monitoring are determining
whether released individuals survive and establish a breeding population,
242 Chapter 7

then following that population over time to see whether it increases in


numbers of individuals and geographic range. Monitoring of important
ecosystem elements is also needed to determine the broader impact of a
reintroduction; for example, when a predator species is introduced, it will
be crucial to determine its effect on prey species and competing species and
its indirect effect on vegetation (Baker et al. 2016). In an otter reintroduc-
tion program, for instance, the returned otter populations appealed to the
general public, but the otters reduced populations of fish and crustaceans,
which angered commercial fishermen (Fanshawe et al. 2003).
The costs of reintroduction also need to be tracked and published so
it can be determined whether reintroduction represents a cost-effective
strategy. In the case of the orangutan (Pongo spp., Figure 7.3), it was found
that reintroduction costs twelve times as much per animal as the protec-
tion of forest habitat. Reintroduction was effective in the short term, but at
time scales longer than 10–20 years, habitat conservation was much more
cost-effective (Wilson et al. 2014).

Behavioral ecology of released animals


To be successful, both introduction and reintroduction programs must often
address the behaviors of animals that are being released (Buchholz 2007).
Behavioral ecology is the study of an animal’s behavior in the context of its
environment and considers the adaptive significance of those behaviors.
When social animals, which include many mammals
and some bird species, grow up in the wild, they learn
from other members of their population, particularly
their parents, how to interact with their environment
and with other members of their species. They learn
how to search for food and how to gather, capture, and
consume it. When mammals and birds are raised in
captivity, their environment is limited to a cage or pen,
so exploration is unnecessary. Searching for food and
learning about new food sources is not required be-
cause the same food items come to them day after day,
on schedule. For example, when the European bison
(Bison bonasus, see chapter opening photo), previously
extinct in the wild, was reintroduced, it was necessary
to determine whether the animals were successfully

Figure 7.3 The Sumatran orangutan (Pongo abelii) is


a critically endangered species found only on the island of
Sumatra, in Indonesia. It is rarer than the other species of
orangutan, the Bornean orangutan. Although reintroduc-
tion programs for both species are common and achieve
instant results, it has been argued that protecting their
habitat may protect more individuals in the long term.
(© Ryan Deboodt/Getty Images.)
Bringing Species Back from the Brink 243

exploring and using their new, wild environment (Schmitz et al. 2015). In
a recent review of conservation projects that use behavioral ecology, it was
found that while foraging and dispersal behaviors were often considered,
others such as anti-predator behavior and social behaviors were not, even
though these are important issues for reintroductions (Berger-Tal et al. 2016).
Social behavior may become highly distorted when animals are raised
alone or in unnatural social groupings (i.e., in small groups or single-aged
groups). In such cases, the animals may lack the skills to survive
in their natural environment and the social skills necessary to
cooperatively find food, sense danger, find mating partners, and
It is imperative that
raise young (Parlato and Armstrong 2013). The greatest threats
captive-bred mammals
to the survival of these animals, and the primary reasons many
and birds learn predator
such establishment projects fail, are predation (“Why are these
avoidance and species-
guys trying to eat me?”), starvation (“Why aren’t they feeding
me any more?”), and habitat quality (“My old home was way appropriate social
better!”) (White et al. 2012). behavior if they are to
To overcome these behavioral problems, captive-raised survive and reproduce
mammals and birds may require extensive training before and after being released into
after release. In some cases, human trainers use puppets or wear the wild. They may also
costumes to mimic the appearance and behavior of wild indi- require some support
viduals so that young animals learn to identify with their own after release.
species rather than with humans (Figure 7.4).
In other cases, wild individuals serve as “in-
structors” for captive individuals of the same
species. For example, wild golden lion tamarins
(Leontopithecus rosalia) are caught and held with
captive-bred tamarins so that the captive-bred
tamarins will learn appropriate behavior from the
wild ones. After they form social groups, they are
released together. Wild-caught African wild dogs
(Lycaon pictus) that gave birth and bonded together
in holding areas prior to release had a higher suc-
cess rate after reintroduction in multiple sites in
South Africa (Gusset et al. 2008).

Establishing plant populations


Methods used to establish new populations of rare
and endangered plant species are fundamentally

Figure 7.4 California condor chicks (Gymnogyps


californianus) raised in captivity are fed by researchers
using puppets that look like adult birds. Conservation
biologists hope that minimizing human contact with
the birds will improve their chances of survival when
they are returned to the wild. (Photograph by Ron Gar-
rison, courtesy of US Fish and Wildlife Service.)
244 Chapter 7

different from those used to establish terrestrial vertebrate species. Ani-


mals can disperse to new locations and actively seek out the most suitable
microsites. The seeds of plants, however, are dispersed to new sites by
agents such as wind, animals, water, or the actions of conservation biolo-
gists (e.g., Rood et al. 2015). Once a seed lands on the ground or an adult is
planted at a site, it is unable to move, even if a suitable microsite exists just
a few meters away. The immediate microsite is crucial for plant survival:
if the environmental conditions are in any way too sunny, too shady, too
wet, or too dry, the seed will not germinate, or the resulting plant will not
reproduce or will die.
Plant ecologists are investigating the effectiveness of site treatments be-
fore and after planting—such as burning the leaf litter, removing compet-
ing vegetation, digging up the ground, and excluding grazing
animals—as a means of enhancing population establishment
New plant populations (Reiter et al. 2016). They are also investigating the most effective
are established by sow- way to establish a species: whether seeds should be sown or
ing seeds or transplant- adult individuals or seedlings should be transplanted into the
ing seedlings or adults. enhanced site. For example, during the introduction of Mead’s
Site treatments such as milkweed (Asclepias meadii), a threatened perennial prairie plant
burning off or physically in the midwestern United States, it was found that survivorship
removing competing was greater for older juvenile plants than for seedlings, and that
plants are often necess- survival was higher in burned habitat than in unburned habitat.
ary for success. Seedling survival was also higher in 1996, which had greater
rainfall than average. Researchers also use PVA (see Chapter 6)
and other modeling tools to determine best practices for plant
reintroduction (Halsey et al. 2015) (Figure 7.5).
Plant reintroductions frequently fail, although, as with animals, this
fact is not captured in the literature because researchers do not usually
publish failed experiments (Godefroid et al. 2011a). Furthermore, many
apparently successful reintroductions have either failed after additional
years or have never established a second generation of plants, which is
an important indicator of success. In general, success increases with the
numbers of individuals or populations reintroduced (Liu et al. 2015) (see
Figure 7.5), but this conclusion assumes that plantings are taking place in
suitable habitat. Genetic analysis of source populations can assist in deter-
mining resistance to environmental stress, and resilience to environmental
change (He et al. 2016). Because of climate change, certain plant species
may no longer be genetically suited to their present sites, and conservation
biologists may have to look elsewhere in the range of a species for suitable
genotypes to plant (Gray et al. 2011).

The status of new populations


The establishment of new populations raises some novel issues at the inter-
section of scientific research, conservation efforts, government regulation,
and ethics. These issues need to be addressed because reintroduction, in-
troduction, and reinforcement programs will increase in the coming years
as the biodiversity crisis eliminates more species and populations from the
Bringing Species Back from the Brink 245

1 population
added
1.0

0.9

0.8

0.7
Probability of extinction

0.6

0.5

0.4

0.3

0.2

0.1 100 populations


added
0
0 10 20 30 40 50
Time since 2012 (years)

Figure 7.5 Metapopulation viability analysis (a modeling tool similar to PVA


that considers number of populations rather than individuals) of Pitcher’s thistle
(Cirsium pitcheri; inset), a federally threatened species that grows in the Indiana
Dunes. The graph shows that the probability of extinction over time was greatly
reduced by increasing numbers of reestablished populations. After 40 years, the
extinction probability is almost 100% with one population added, but is reduced
to less than 1% with more than 100 populations added. (After Halsey et al. 2015.
Photograph courtesy of the US Fish and Wildlife Service.)

wild. In addition, assisted colonization may be needed for many species


if their present ranges become too hot, too dry, or otherwise unsuitable
because of global climate change or some other change in the environment
(Lunt et al. 2013).
Many of the reintroduction programs for endangered species are man-
dated by official recovery plans set up by national governments. If such
plans are to be formulated and implemented, conservation biologists must
be able to explain the benefits and limitations of reintroduction and in-
troduction programs in a way that government officials and the general
public can understand, and they must address the legitimate concerns of
those groups (Figure 7.6).
Sometimes stakeholders’ concerns can be addressed by giving vary-
ing degrees of protection to new populations. In the United States, for ex-
ample, populations can be designated experimental essential or experimental
nonessential. Experimental essential populations are regarded by the US
Endangered Species Act as critical to the survival of the species, and they
246 Chapter 7

Figure 7.6 Residents of southern Taiwan were


surveyed to assess attitudes toward a restoration pro-
gram for Formosan sika deer (Cervus nippon taiouanus),
which had gone extinct in the wild, in Kenting National
Park. After its reintroduction, the deer was associated
with threats to crops (48% of respondents), and 18% of
respondents had actually suffered losses to the deer.
Even so, the majority of those surveyed believed the
deer was a tourism resource (87%) and supported
the reintroduction program (75%). (Yen et al. 2015.)
(© Imagemore Co., Ltd./Getty Images.)

are as rigidly protected as naturally occurring popu-


lations. Experimental nonessential populations have
less protection under the law; designating popula-
tions as nonessential often helps to overcome the fear
of local landowners that having endangered species
on their property will restrict how their land can be
managed and developed.
Legislators and scientists alike must understand
that the establishment of new populations through
reintroduction programs in no way reduces the need
to protect the original populations of the endangered
species. Original populations are likely to have the
most complete gene pool of the species and the most
intact interactions with other members of the biological community. In
many cases, proposals are made by developers or government departments
to compensate for habitat damage or eradication of endangered
populations by development projects by creating new habitat or
establishing new populations. This activity is generally referred
The establishment of
to as mitigation. Mitigation plans, which are often directed at
new populations through
legally protected species and habitats, often include (1) adjust-
reintroduction programs
ments to the development plan to reduce the extent of damage,
in no way reduces the
(2) establishment of new populations and habitat as compensa-
need to protect the
tion for what is being destroyed, and (3) enhancement of popu-
original populations of lations and habitat that remain after development. Given the
endangered species. poor success of most attempts to create new populations of rare
species, mitigation plans for threatened or endangered species
should be approached with great skepticism.

Ex Situ Conservation Strategies


The best strategy for the long-term protection of biodiversity is the preserva-
tion of biological communities and populations (called in situ conservation,
as discussed in Chapter 6). However, in the face of increasing threats to
biodiversity, relying solely on in situ conservation is not currently a viable
Bringing Species Back from the Brink 247

option for many rare and endangered species. For example, even under in
situ conservation management and protection programs, species may still
decline and go extinct in the wild for any of the reasons already discussed:
habitat destruction, loss of genetic variation, demographic and environmen-
tal stochasticity, and so forth. Likewise, if a remnant population is too small
to maintain the species, if it is still declining despite conservation efforts, or
if the remaining individuals are found outside of protected areas, then in
situ conservation may not be adequate. It is likely that the only
way to prevent species in such circumstances from going extinct
is to maintain individuals in artificial conditions under human When integrated with
supervision (Canessa et al. 2015). Ex situ, or off-site, conservation efforts to protect existing
used in place of, or to complement, in situ conservation can mean populations and to
the difference between persistence and extinction for some spe- establish new ones, ex
cies. Already a number of species that have gone extinct in the situ conservation is an
wild have survived because of propagation in captive colonies. important strategy for
The beautiful Franklin tree (Franklinia alatamaha), for example, protecting endangered
grows only in cultivation and is no longer found in the wild. species and educating
Ex situ and in situ conservation are complementary strate- the public.
gies (Zimmermann et al. 2007) (Figure 7.7). The long-term goal
of many ex situ conservation programs is the establishment of
new populations in the wild, once sufficient numbers of indi-
viduals and a suitable habitat are available. In the case of Prze- Figure 7.7 This model
walski’s horse (Equus ferus przewalskii), which had been declared shows ways in which in situ (on-
extinct in the wild, small herds descended from 14 captive-bred site) and ex situ (off-site) conser-
vation efforts can benefit each
other and provide alternative
ENDANGERED SPECIES conservation strategies. While
no species conforms exactly to
this idealized model, the giant
Field survey panda program (see Figure 7.11)
has many of its elements. (After
Maxted 2001.)
Conservation strategies

EX SITU CONSERVATION IN SITU CONSERVATION

Revenues fund
Storage of sperm, eggs, Zoos, aquariums, conservation efforts Protection, management,
tissue, seeds, etc. botanical gardens monitoring
Collections
from the wild
Establishing breeding Viable wild populations
programs Reestablish
populations
in the wild
Develop new products
Funds to maintain Funds to protect
breeding programs and manage species
Use and sell new products in the wild
248 Chapter 7

founder individuals were released in a national park in Mongolia starting


in 1992 (Figure 7.8A).
An intermediate strategy that combines elements of both ex situ and in
situ conservation is the monitoring and management of populations of rare
and endangered species in small protected areas. Such populations are still
somewhat wild, but occasional human intervention may be necessary to
prevent population decline. An example is the rare Columbia Basin pygmy
rabbit (Brachylagus idahoensis), which does not breed well in captivity (Fig-
ure 7.8B). As part of the conservation strategy for this species, groups of
pygmy rabbits are released into large (2.5 ha) fenced enclosures that keep
out predators, keep the rabbits in familiar surroundings, and encourage
breeding. On occasion, tagged individuals from the breeding program are
released back into the wild.
Research on captive populations can provide insight into the basic biol-
ogy, physiology, and genetics of a species through studies that would not
be possible on wild animals. The results
of these studies can suggest new conser-
(A)
vation strategies for in situ populations.
Long-term, self-sustaining ex situ popula-
tions can also reduce the need to collect
individuals from the wild for display
or research. Ex situ facilities for animal
preservation include zoos, game farms,
and aquariums as well as the facilities of
private breeders. Plants are maintained
in botanical gardens, arboretums, and
seed banks.
Ex situ conservation has several limi-
tations. First, it is not cheap. The cost of
maintaining zoos is enormous in com-
parison with many other conservation
(B)

Figure 7.8 (A) The IUCN has declared


Przewalski’s horse to be extinct in the wild.
Several zoos around the world have main-
tained populations and have been success-
ful in breeding these animals. Reintroduc-
tions of the wild horses into their natural
range in Mongolia appear so far to have
been successful. (B) Rare Columbia Basin
pygmy rabbits breed well in large seminatu-
ral enclosures. Individuals are then released
into the wild. (A, Patrick Pleul/dpa/Corbis;
B, photograph by Tara Davila, courtesy of
Lisa Shipley and Rod Sayler, Washington
State University.)
Bringing Species Back from the Brink 249

activities, such as the budgets of national parks in developing countries;


in the United States alone, zoos cost about $1 billion per year to run. Fur-
thermore, ex situ programs protect only one species at a time. In contrast,
when a species is preserved in the wild, an entire community—perhaps
consisting of thousands of species—may be preserved, along with a range
of ecosystem services.
There are also several ecological and evolutionary problems with ex
situ conservation. Only in natural biological communities are species able
to continue their process of evolutionary adaptation (Olivieri et al. 2016).
Ecosystem-level interactions among species, as discussed in Chapter 2, are
often crucial to a rare species’ continued survival; these interactions can be
quite complex and probably cannot be replicated under captive conditions.
Furthermore, captive animal populations are generally not large enough
to prevent the loss of genetic variation through genetic drift; the same may
be true of cultivated plant species when they have special requirements for
pollination that might make it difficult to ensure adequate cross-fertiliza-
tion among individuals. For such species, the best solution may be in situ
conservation involving careful habitat protection and management. Despite
these problems, ex situ conservation remains an important approach, and
many species have recovered from near extinction.
Captive individuals on display can also serve as ambassadors for their
species and help to educate the public about the need to preserve the spe-
cies in the wild (Figure 7.9A). Zoos, aquariums, botanical gardens, and
the people who visit them regularly contribute money and expertise to in
situ conservation programs (see Figure 6.2). In addition, ex situ conserva-
tion programs can be used to develop new products that can potentially
generate funds from profits or licensing fees to protect species in the wild.
In situ preservation of species, in turn, is vital to the survival of species
that are difficult to maintain in captivity, as well as to the continued ability
of zoos, aquariums, and botanical gardens to display species that do not
have self-sustaining ex situ populations.

Zoos
A current goal of most major zoos is to establish viable, long-term captive
breeding populations of rare and endangered animals (Zimmermann et
al. 2008). The International Zoo Yearbook (IZY) reports births and deaths of
zoo-bred animals, and a review of these data since 1972 suggests that all but
five endangered mammalian species have median positive captive popula-
tion growth rates, even if those rates have decreased over time (Alroy 2015).
Zoos, along with affiliated universities, government wildlife departments,
and conservation organizations, presently maintain over 2 million animals,
including over 600,000 individual terrestrial vertebrates, representing over
7400 species and subspecies of mammals, birds, reptiles, and amphibians
(Table 7.1) (www2.isis.org). While this number of captive animals may
seem impressive, it is trivial in comparison to the tens of millions of domestic
cats, dogs, and fish kept by people as pets. Zoos could establish breeding
250 Chapter 7

TABLE 7.1 Number of Terrestrial Vertebrates Maintained in Zoos


Location Mammals Birds Reptiles Amphibians Total
Europe 101,921 125,846 30,799 57,413 315,979
North America 50,982 62,448 31,270 50,588 195,288
Latin America 3,653 5,105 2,455 634 11,847
Asia 29,089 39,216 9,338 887 78,530
Australasia 7,674 10,312 3,890 1,875 23,751
Africa 4,185 7,939 2,435 356 14,915
Worldwide totals
All species 197,504 250,866 80,187 111,753 640,310
Number of taxaa 2,238 3,753 969 544 7,486
Percentage wild-bornc 5% 9% 15% 5%
Rare speciesb 59,030 37,748 22,474 3,398 122,650
a
Number of taxa 527 344 207 29 1,107
Percentage wild-bornc 7% 9% 18% 7%
Source: Data from ISIS, provided by Laurie Bingaman Lackey 2013.
a
The number of taxa is not exactly equivalent to number of species because many species have more
than one subspecies listed.
b
Rare species are those covered by CITES (the Convention on International Trade in Endangered
Species).
c
The percentage of individuals born in the wild is approximate (particularly for reptiles and amphib-
ians), since the origin of the animals is often not reported.

colonies of even more species if they directed more of their efforts toward
smaller-bodied species such as insects, amphibians, and reptiles, which are
less expensive to maintain in large numbers than are large-bodied mammals
such as bears, elephants, and rhinoceroses. A better balance must be reached
between displaying large animals that draw many visitors and displaying
smaller, lesser-known animals that appeal less to the public but represent a
greater proportion of the world’s biodiversity.
Zoos already work together effectively to conserve some of these small-
er species. For instance, seven North American zoos joined with universi-
ties and the Defenders of Wildlife to form the Panama Amphibian Rescue
and Conservation Project (www.amphibianrescue.org). A major goal of
this collaboration is to establish breeding populations of frogs and other
amphibians that are being decimated in the wild (Figure 7.9B). Ex situ
conservation efforts have been increasingly directed at saving endangered
species of invertebrates as well, including butterflies, beetles, dragonflies,
spiders, and molluscs. Other important targets for ex situ conservation
efforts are rare breeds of domestic animals on which human societies
depend for animal protein, dairy products, leather, wool, agricultural labor,
transport, and recreation (Ruane 2000). Secure populations of these breeds
are a potential genetic resource for the improvement and long-term health
of our supplies of pigs, cattle, chickens, sheep, and other domestic animals.
Bringing Species Back from the Brink 251

(A)

(B)
550

500

450
Number of species

400

350

300
Total non-GTS
250
Total VU species
200 Total EN species
150 Total CR and EW species

100
1994 1996 1998 2000 2002 2004 2006 2008 2010 2012 2014
Year

Figure 7.9 (A) Mother with her young daughter watch an aquarist wearing
a wetsuit gently hold a Chinese giant salamander in a freshwater tank. Criti-
cally endangered in the wild, the Chinese giant salamander is the largest living
species of amphibian. (B) The number of amphibian species at zoos has been
steadily increasing since 1994, not only for non–globally threatened species
(non-GTS), but also for vulnerable (VU), endangered (EN) and critically endan-
gered (CR) species, and species extinct in the wild (EW). (A, photograph © Billy
Hustace/Corbis; B, after Dawson et al. 2016.)

Introduction to Conservation Biology 1E Primack/Sher


252 Chapter 7

The variety of species displayed in zoos has increased in recent years,


but the emphasis is still on “charismatic megafauna” because they help to
attract the general public and influence them favorably toward conserva-
tion. In fact, over 90% of families enjoy seeing biodiversity at zoos and
aquariums and believe they teach children about protecting species and
habitat (www.aza.org). As such, ex situ animals make people aware of the
threats to biodiversity (Moss et al. 2016). As part of the World Zoo Con-
servation Strategy, which seeks to link zoo programs with conservation
efforts in the wild, the world’s 2000 zoos and aquariums are increasingly
incorporating ecological themes and information about the threats to en-
dangered species into their public displays and research programs. The
potential educational and financial impacts of zoos are enormous, consid-
ering that they receive approximately
600 million visitors per year (Figure
(A) 7.10A). Furthermore, the funds raised
by zoos from visitor fees and other
programs can directly fund in situ
conservation activities.
Zoos have the needed knowledge
and experience in animal care, vet-
erinary medicine, animal behavior,
reproductive biology, and genetics to
establish captive animal populations
of endangered species (Figure 7.10B)
(Zimmermann et al. 2008). Zoos and
affiliated conservation organizations
have embarked on major efforts to
build the facilities and develop the
technology necessary to establish and
house breeding colonies of endangered
animals and to develop the new meth-
(B) ods and programs needed to reintro-
duce species in the wild (Figure 7.11).

Figure 7.10 (A) Zoos can educate


the public about the need to protect
wildlife. Here, visitors to the Asahiyama
Zoo on Japan’s northern island of Hok-
kaido enjoy a parade of king penguins
(Aptenodytes patagonicus). (B) Veterinar-
ians carry out dental surgery on a captive
Asian elephant. The knowledge gained
by caring for captive animals can be ap-
plied to helping the species in the wild.
(A, Photograph © JTB Photo Communi-
cations, Inc./Alamy; B, © Richard Clem-
ent/Zuma Press.)
Bringing Species Back from the Brink 253

Figure 7.11 China’s giant panda (Ailuropoda


melanoleuca) is one of the world’s most charis-
matic animals and has become emblematic of
the fight to save endangered species. Wolong Na-
tional Nature Reserve and other facilities in China
have been successful at breeding giant pandas
using artificial insemination and hand rearing.
The reserve has established a reintroduction
program, but the loss of habitat and the fact that
captive-raised individuals may lack the behavior-
al skills needed to survive in the wild combine to
make reintroduction of pandas particularly prob-
lematic. (Photograph © LMR Group/Alamy.)

Some of these facilities are highly specialized,


such as that run by the International Crane
Foundation in Wisconsin, which is attempt-
ing to establish captive breeding colonies of all
crane species. These collaborative efforts have
paid off. Currently, fewer than 5% of the terrestrial mammals
kept in zoos have been collected in the wild, and this number is Zoos often use the latest
declining as zoos gain more experience. methods of veterinary
The success of captive breeding programs has been enhanced medicine to establish
by efforts to collect and disseminate knowledge about the main- healthy breeding colonies
tenance of rare and endangered species. The Species Survival of endangered animals.
Commission’s Conservation Breeding Specialist Group, a divi-
sion of the IUCN, and affiliated organizations, such as the As-
sociation of Zoos and Aquariums, the European Association of
Zoos and Aquaria, and the Australasian Regional Association of Zoological
Parks and Aquaria, provide zoos with the necessary information on proper
care and handling of these species as well as updates on the status and be-
havior of the animals in the wild (www.aza.org). This information includes
nutritional requirements, anesthetic techniques to immobilize animals and
reduce stress during transport and medical procedures, optimal housing
conditions, information about vaccinations and antibiotics to prevent the
spread of disease, and breeding records. The collective effort is being aided
by a central database called the Zoological Information Management System
(ZIMS), maintained by the International Species Information System (ISIS),
which keeps track of all relevant information on the over 2 million animals
belonging to 10,000 vertebrate and invertebrate species at 825 member in-
stitutions more than 85 countries (www2.isis.org).
Many species provided with humane captive conditions reproduce with
abandon—so much so that the use of contraceptives to control populations
is required. However, some rare animal species, such as rhinoceroses, do
not adapt or reproduce well in captivity (McCleery et al. 2014). Zoos con-
duct extensive research to identify management methods that can over-
254 Chapter 7

come these problems and promote successful reproduction of genetically


appropriate mates (Wildt et al. 2009). Some of these methods come directly
from human and veterinary medicine, while others are novel techniques
developed at special research facilities such as the Smithsonian Conser-
vation Biology Institute in Virginia and the San Diego Zoo Institute for
Conservation Research. These techniques include cross-fostering, in which
common species raise the offspring of rare species; artificial incubation of
eggs under ideal hatching conditions; artificial insemination when adults
do not show interest in mating or are living in different locations; and
embryo transfer, which involves implanting fertilized eggs of rare species
into surrogate mothers of common species (Figure 7.12). One of the most
unusual techniques, known as genome resource banking (GRB), involves
the freezing of purified DNA, eggs, sperm, embryos, and other tissues
so that they can be used to contribute to breeding programs, to maintain
genetic diversity, and for scientific research. However, many of these tech-
niques are expensive and species specific. In any case, GRB and similar
advanced methods are not substitutes for in situ and ex situ conservation
programs that preserve behaviors and ecological relationships that are
necessary for survival in the wild.
Raising animals in captivity places limits on the animals’ conservation
value. Animals living in captive breeding facilities may lose the behaviors
they need to survive in the wild. Furthermore, populations raised in captiv-
ity may undergo genetic, physiological, and morphological changes that

Figure 7.12 A Brazilian ocelot


produced by embryo transfer at the
Cincinnati Zoo. (Courtesy of the
Cincinnati Zoo.)
Bringing Species Back from the Brink 255

make them less able to tolerate the natural environment if they are returned
to the wild. Diseases acquired in captivity may render them unsuitable
for release (Minuzzi-Souza et al. 2016). Consequently, when researchers
establish an ex situ program to preserve a species, they must address a
series of ethical questions:
• How will establishing an ex situ population benefit the wild population?
• Is it better to let the last few individuals of a species live out their days
in the wild or to breed a captive population that may be unable to
adapt to wild conditions?
• Does a population of a rare species consisting of individuals that have
been raised in captivity and do not know how to survive in their natu-
ral environment really represent preservation of the species?
• Are rare individuals being held in captivity primarily for their own
benefit, for the benefit of their entire species, for the economic benefit
of zoos, or for the pleasure of zoo visitors?
• Are the animals in captivity receiving appropriate care based on their
biological needs?
• Are sufficient efforts being made to educate the public about conser-
vation issues?

Aquariums
Approximately 600,000 individual fish, most of them obtained from the
wild, are maintained in public aquariums that are open to visitors (Figure
7.13). Major efforts are being made to develop breeding techniques so that
rare species can be maintained in aquariums without further collection
in the wild and in the hope that some can be released back into the wild.

Figure 7.13 Public aquari-


ums participate in both in situ
and ex situ conservation pro-
grams and provide a valuable
function by educating people
about marine conservation is-
sues. (© Jeff Greenberg/Alamy
Stock Photo.)
256 Chapter 7

Fish breeding programs use indoor aquarium facilities, seminatural


water bodies, and fish hatcheries and farms. Many fish breeding tech-
niques were originally developed by fisheries biologists for large-scale
stocking operations involving trout, bass, salmon, and other commercial
species. Other techniques were discovered in the aquarium pet trade when
dealers attempted to propagate tropical fish for sale. These techniques are
now being applied to endangered freshwater fauna. Currently, both public
and private groups are making impressive efforts to unlock the secrets of
propagating some of the more difficult species.
Aquariums have an increasingly important role to play in the conserva-
tion of endangered cetaceans, manatees, sea turtles, and other large marine
animals. Aquarium personnel often respond to public requests for assistance
in handling large animals stranded on beaches or disoriented in shallow
waters. The aquarium community can use lessons learned from working
with common species to develop programs to aid endangered species.
The ex situ preservation of aquatic biodiversity takes on additional sig-
nificance due to the dramatic recent increase in aquaculture, which repre-
sents about 30% of fish and shellfish production worldwide. This aquaculture
includes the extensive salmon, carp, and catfish farms in the temperate zones,
the shrimp and fish farms in the tropics, and the 12 million tons of aquatic
products grown in China and Japan. As fish, frogs, molluscs, and crustaceans
are increasingly domesticated and raised to meet human needs, it becomes
necessary to preserve the genetic stocks needed to maintain and increase
commercial production of these species and to protect them against disease
and unforeseeable threats. A challenge for the future will be balancing the
need to increase human food production through aquaculture with the need
to protect aquatic biodiversity from increasing human threats.

Botanical gardens
The world’s 1775 botanical gardens (also known as botanic gar-
Botanical gardens have dens) contain major collections of living plants and represent a
living collections and crucial resource for plant conservation through ex situ conser-
seed banks that provide vation, research, and education (Figure 7.14). An arboretum
ex situ protection and is a specialized botanical garden focusing on trees and other
knowledge of endangered woody plants. The world’s botanical gardens currently contain
and economically about 4 million living plants, representing 80,000 species—ap-
important plants. proximately 30% of the world’s flora (Guerrant et al. 2013; www.
bgci.org). When we add in the species grown in greenhouses,
subsistence gardens, and hobby gardens, the numbers are in-
creased. One of the world’s largest botanical gardens, the Royal
Botanic Gardens, Kew, in England, has over 30,000 species of plants under
cultivation, about 10% of the world’s total; 2700 of these species are listed as
threatened by the IUCN. One of the most exciting new botanical gardens is
the Eden Project in southwestern England, which focuses on displaying and
explaining over 5000 species of rain forest, temperate-zone, and Mediter-
ranean plants in giant domes that constitute the world’s largest greenhouse
(www.edenproject.com).
Bringing Species Back from the Brink 257

(A) (B)

(C) (D)

Figure 7.14 (A) The beautiful displays at Munich Botanical Gardens in


Germany play an important role in conservation by connecting the public
with plants and preserving those species that no longer occur in the wild. (B)
Signage at the Kirstenbosch National Botanical Garden, South Africa, edu-
cates visitors about how to promote local biodiversity in their yards. (C) Field
biologists with Madagascar program of the Missouri Botanical Garden identify
plants that may have medicinal benefits, work to safeguard the biodiverse ar-
eas where they grow, and raise plants for reintroductions. (D) A scientist at the
Royal Botanical Gardens, Kew, United Kingdom inspects preserved seeds as a
part of the Millennium Seed Project. (A, courtesy of Richard Primack; B, cour-
tesy of Anna Sher; C, courtesy of Missouri Botanical Gardens; D, courtesy of the
Royal Botanic Gardens, Kew.)

Botanical gardens increasingly focus their efforts on cultivating rare


and endangered plant species, and many specialize in particular types
of plants. Many botanical gardens are involved in plant conservation, es-
pecially in the reintroduction of rare and endangered plant species and
the restoration of degraded ecosystems (Hardwick et al. 2011). In Yunnan,
China, only 52 individuals of a rare magnolia tree, Magnolia sinica, are left
258 Chapter 7

in the wild, but the species has been successfully cultivated at Kunming
Botanical Garden, providing hope that reintroductions may be possible.
Staff members at botanical gardens are often recognized authorities
on plant identification, distributions, and conservation status. Botanical
gardens are able to educate an estimated 200 million visitors per year about
conservation issues. At an international level, Botanic Gardens Conser-
vation International (BGCI) represents and coordinates the conservation
efforts of over 700 botanical gardens (www.bgci.org). The priorities of this
program include the creation of a worldwide database to support collecting
activity and identification of important species that are underrepresented
or absent from collections of living plants. One of its projects is the on-
line PlantSearch database, which currently lists over 1.3 million records
of more than 480,000 species and varieties growing in botanical gardens,
of which about 3000 are rare or threatened. In addition, because most
existing botanical gardens are located in the temperate zone, establishing
botanical gardens in the tropics is a primary goal of the international
botanical community.

Seed banks
In addition to growing plants, botanical gardens and research institutes
have developed collections of seeds, sometimes known as seed banks,
obtained from the wild and from cultivated plants. These seed banks pro-
vide a crucial backup to their living collections (Figure 7.15). The seeds
of most plant species can be kept dormant in cold, dry conditions for long
periods and later germinated to produce new plants. This ability of seeds
to remain dormant allows the seeds of large numbers of rare species to
be frozen and stored in a small space, with minimal supervision and at
a low cost. Seed banks are especially important for rare and endangered
species that may need to be reintroduced into the wild. At present, seeds
of approximately 4000 or 10% of the world’s known plant species are stored
in seed banks, with the figure approaching 70% for European plants (Go-
defroid et al. 2011b). Efforts are made to include the full range of genetic
variation found in a species by collecting seeds from populations growing
across the range of the species.
More than 1000 seed banks exist worldwide, and their activities are
coordinated by the Consultative Group on International Agricultural
Research (CGIAR). However, if power supplies fail or equipment breaks
down, an entire frozen collection could be damaged. To prevent such a
loss, Norway has recently established the newest seed bank, the Svalbard
Global Seed Vault, where 400,000 frozen seed samples are stored below
permafrost, with millions more expected to be added in coming decades.
Seed banks have been embraced by agricultural research institutes and
the agricultural industry as an effective resource for preserving and using
the genetic variation that exists in agricultural crops and their wild relatives.
Preserving this genetic diversity is crucial to maintaining and increasing the
high productivity of modern crops and their ability to respond to changing
Bringing Species Back from the Brink 259

(A) (B)

Figure 7.15 (A) At seed banks, seeds of many plant varieties are sorted,
cataloged, and stored at freezing temperatures. (B) Seeds come in a wide variety
of sizes and shapes. Each such seed represents a genetically unique, dormant in-
dividual. (Photographs courtesy of US Department of Agriculture.)

environmental conditions such as acid rain, global climate change, and soil
erosion (Banga and Kang 2014). Researchers are in a race against time to
preserve genetic variation because traditional farmers throughout the world
are abandoning their diverse local crop varieties in favor of standard, high-
yielding varieties (Gliessman 2015). By some estimates, 75% of global crop
plant genetic diversity has been lost in this way over the last century (FAO
2007). This worldwide phenomenon is illustrated by Sri Lankan farmers,
who grew two thousand varieties of rice until the late 1950s, when they
switched over to just five high-yielding varieties.
A major controversy involved in the development of agricultural seed
banks is who owns and controls the genetic resources of crops (Brush
2007). The genes of landraces of crop plants (local species that have been
adapted by humans over time) and wild relatives of crop species represent
the building blocks needed to develop elite, high-yielding varieties suit-
able for modern agriculture (Nabhan 2008). Approximately 96% of the raw
genetic variation necessary for modern agriculture comes from develop-
ing countries such as India, Ethiopia, Peru, Mexico, Indonesia, and China
(Figure 7.16), yet most corporate breeding programs for elite strains are
located in the industrialized countries of North America and Europe. In
the past, genetic material was perceived as free for the taking: the staffs
of international seed banks freely collected seeds and plant tissue from
developing countries and gave them to research stations and seed compa-
260 Chapter 7

Soybean
Grapes

Wheat

Barley
Cabbage,
lettuce,
olives, oats

Bean
Rubber, Rice
chocolate

Citrus
Tomato
Corn

Coffee, Sugarcane
sorghum,
wheat
Cotton
Banana,
Potato Peanut yam, coconut

Figure 7.16 Crop species show high genetic diversity in certain areas of
the world (shown in blue), often where the species was first domesticated or
where it is still grown in traditional agricultural settings. This genetic diversity is
of international importance in maintaining the productivity of agricultural crops.
(Map courtesy of Garrison Wilkes.)

nies. Seed companies then developed new strains through sophisticated


breeding programs and field trials. The resulting seeds were sold at high
prices to maximize profits for the companies, which often totaled hundreds
of millions of dollars a year, but the countries from which the original seeds
were collected did not receive any profit from this activity. Developing
countries now question why they should share their biological materials
freely if they will have to pay for new seed varieties and cultivated plants
based on those genetic resources.
One solution to this controversy involves negotiating agreements, using
the framework of the Convention on Biological Diversity (see Chapter 11),
in which countries agree to share their genetic resources in exchange for
receiving new products and a share of the profits. A few such contracts
have been negotiated, such as the one between Merck and the government
of Costa Rica for the development of products based on species collected
from the wild in that country (see Chapter 3 for similar examples). These
Bringing Species Back from the Brink 261

agreements will be followed carefully to determine whether they are mutu-


ally satisfactory and can serve as models for future contracts.

Can Technology Bring Back Extinct Species?


The idea of using technology to bring species back from extinction is not
particularly new. Medical scientists have re-created viruses and other
pathogens in the past to study their behavior in hope of preventing simi-
lar diseases. In 2003, scientists even brought back an extinct mammal: the
Pyrenean ibex or bucardo (Capra pyrenaica pyrenaica), the last of which had
died in 2000. The process involved fusing frozen bucardo cells with goat
egg cells from which the nuclei had been removed and implanting the
resulting eggs into a domestic goat, which served as a surrogate mother.
Even though an embryo did develop, the resulting cloned ibex was badly
deformed and died within minutes of birth.
The process of bringing a species back—sometimes referred to as de-
extinction—requires DNA, ideally a complete genome. This requirement
limits the candidate species to those that have not been extinct so long
that their genetic material has completely degraded. The genome of the
extinct species, perhaps obtained from tissue preserved in a freezer, bur-
ied under a glacier, or preserved deep in an anoxic bog or oceanic mud, is
then compared with that of a closely related living species—for example,
the extinct passenger pigeon with the extant band-tailed pigeon—and the
differences identified. At that point, the living species’ DNA and cells can
potentially be modified and used to create stem cells and germ cells with
the DNA of the extinct species. These germ cells can then be implanted
into eggs from living relatives. At the end of the gestation period, the
result could be an extinct species brought back to life. This process was
recently used to create early-stage embryos of an extinct gastric-brooding
frog (Rheobatrachus silus), but they soon died. Perhaps in the future, the
technique could be applied to tissue from extinct species preserved in
the northern permafrost, such as the extinct mastodon, or even to bits of
DNA from fossilized bones.
But is it a good idea to bring back extinct species? The possibility brings
up questions of ethics, practicality, and the fundamental goals of conserva-
tion. For example, if re-created passenger pigeons were brought up by an-
other pigeon species and lacked the flocking, feeding, and other behaviors
of extinct passenger pigeons, should they really be considered passenger
pigeons? Would resurrected passenger pigeons function in an ecosystem
in the same ways as past passenger pigeons? Would it be acceptable to
bring back a species if its habitat no longer exists?
It is also unclear how de-extinction programs would affect and interact
with existing conservation efforts. Would the ability to bring species back
from extinction reduce the incentive to preserve species in the wild? Will
de-extinction research and efforts divert funding from research and con-
servation projects aimed at preserving existing species and ecosystems?
262 Chapter 7

The thought-provoking questions raised by these exciting new tech-


niques highlight the interdisciplinary nature of conservation biology—its
liberal overlap with biotechnology, genetics, ecology, ethics, and econom-
ics. Just because we can do something does not necessarily mean that we
should; whether or not we use de-extinction to bring back an extinct species
or restore its ecosystem functions will depend on the development of new
conservation technologies and the thoughtful discussions that they inspire.

Summary
„„New populations of rare and endan- „„Some species that are in danger of going
gered species can be established in the extinct in the wild can be maintained in
wild using either captive-raised or wild- zoos, aquariums, botanical gardens, and
caught individuals. seed banks; this strategy is known as ex
„„Long-term monitoring of a species in the situ conservation. These captive colonies
field can determine if it is stable, increas- can sometimes be used later to establish
ing, or declining in abundance over new populations in the wild.
time. „„Technologies are being developed that
„„Animals sometimes require behavioral could potentially revive versions of
training before release as well as mainte- extinct organisms, however this may
nance after release. neither be practical nor advisable.

For Discussion
1. How do you judge whether a reintroduc- as described by Donlan et al. (2006)?
tion project is successful? Develop simple What would be some of the legal, eco-
and then increasingly detailed criteria to nomic, and ecological issues involved?
evaluate a project’s success. 3. Would biodiversity be adequately pro-
2. Would it be a good idea to create new tected if every species were raised in
wild populations of African rhinoceroses, captivity? Is this possible? Is it practical?
elephants, and lions in Australia, South How would freezing a tissue sample of
America, the southwestern United States, every species help to protect biodiver-
or other areas outside their current range, sity? Again, is this possible or practical?

Suggested Readings
Banga, S. S. and M. S. Kang. 2014. Developing climate-resilient crops. Journal of
Crop Improvement 28: 57–87. Farmers are crucial to maintaining the genetic
diversity of crops in the face of climate change.
Dawson, J., F. Patel, R. A. Griffiths, and R. P Young, R. P. 2016. Assessing the
global zoo response to the amphibian crisis through 20-year trends in cap-
tive collections. Conservation Biology 30(1): 82–91. Zoos are increasing num-
bers of endangered amphibians in ex situ.
Bringing Species Back from the Brink 263

Dolman, P. M., N. J. Collar, K. M. Scotland, and R. J. Burnside. 2015. Ark or


park: stochastic population modeling to evaluate potential effectiveness of
in situ and ex situ conservation for a critically endangered bustard. Journal
of Applied Ecology 52: 841–850. Mathematical models can inform conserva-
tion approaches with less risk to extant populations.
He, X., M. L. Johansson, and D. D. Heath. 2016. Role of genomics and tran-
scriptomics in selection of reintroduction source populations. Conservation
Biology. Genetic analysis of wild populations.
Liu, H., H. Ren, Q. Liu, X. Wen, M. Maunder, and J. Gao. 2015. Translocation
of threatened plants as a conservation measure in China. Conservation Biol-
ogy, 29(6): 1537–1551. Success of rare plant reintroductions in China were
influenced by the number planted, the plant life form, and the source of
the material.
McCleery, R., J. A. Hostetler, and M. K. Oli. 2014. Better off in the wild? Evalu-
ating a captive breeding and release program for the recovery of an endan-
gered rodent. Biological Conservation 169: 198–205. Such programs will only
be successful if animals breed well in captivity and if released animals
have a reasonable rate of survival and reproduction in the wild.
Miller, B., W. Conway, R. P. Reading, C. Wemmer, D. Wildt, D. Kleiman, and
4 others. 2004. Evaluating the conservation mission of zoos, aquariums,
botanical gardens, and natural history museums. Conservation Biology 18:
86–93. Eight tough questions are asked with hopes that these institutions
can become more effective.
Morrell, V. 2014. Science behind plan to ease wolf protection is flawed, panel
says. Science 343: 719. Removing federal protection for the gray wolf in the
United States is controversial, mixing together science and politics.
Sebastián-González, E., J. A. Sánchez-Zapata, F. Botella, J. Figuerola, F. Hiral-
do, and B. A. Wintle. 2011. Linking cost efficiency evaluation with popu-
lation viability analysis to prioritize wetland bird conservation actions.
Biological Conservation 144: 2354–2361. Different management approaches
are evaluated for their cost effectiveness on bird populations in Spain.
Tlusty, M. F., A. L. Rhyne, L. Kaufman, M. Hutchins, G. M. Reid, C. Andrews,
and 4 others. 2013. Opportunities for public aquariums to increase the
sustainability of the aquatic animal trade. Zoo Biology 32: 1–12. Aquariums
should be taking the lead in conservation issues such as the ornamental
fish trade and the sustainable harvesting and consumption of seafood.
Zimmer, C. 2013. Bringing them back to life: The revival of an extinct species
is no longer a fantasy. But is it a good idea? National Geographic 223(14). A
popular review of the science of de-extinction.
Zimmermann, A., M. Hatchwell, L. Dickie, and C. D. West (eds.) 2008. Zoos
in the 21st Century: Catalysts for Conservation, pp. 1243–1248. Cambridge
University Press, Cambridge. Many modern zoos see the advancement of
in situ wildlife conservation as part of their mission. This often-cited work
suggests that population control is the only solution to the overuse of
common-property resources.
8 Protected Areas
Establishment and Classification Landscape Ecology and Park
of Protected Areas 266 Design 286
Designing Protected Areas 277 Managing Protected Areas 288
Networks of Protected Challenges to Park
Areas 283 Management 298

A designated wilderness area near Gunnison, Colorado.


Wilderness areas receive some of the strongest
protections available.
O
ne of the most used, successful, and surpris-
ingly flexible of the tools for conservation is
the creation of protected areas. A protected
area is a clearly defined geographical space, recog-
nized, dedicated, and managed, through legal or other
effective means, to achieve the long-term conserva-
tion of nature with associated ecosystem services and
cultural values (Dudley 2008). The importance of es-
tablishing protected areas is highlighted by the Aichi
Biodiversity Target 11 from the United Nations Con-
vention on Biological Diversity (see Chapter 11):
By 2020, at least 17 per cent of terrestrial and inland
water areas and 10 per cent of coastal and marine
areas, especially areas of particular importance for
biodiversity and ecosystem services, are conserved
through effectively and equitably managed,
ecologically representative and well-connected
systems of protected areas and other effective area-
based conservation measures, and integrated into
the wider landscape and seascape.
Since passage of the Yosemite Grant Act and the
establishment of the first formal protected area in 1864
and the first national park in 1872 (Figure 8.1), in-
creasing numbers of protected areas have been estab-
lished. However, it is estimated that to reach the goals
for Target 11, an additional 2.2 million km2 are needed
(UNEP Protected Planet Report 2014).
266 Chapter 8

Figure 8.1 American bison (Bison bison) graze at Yellowstone National


Park, Wyoming, the first National Park established in the world. (Photograph by
Richard Primack.)

Protecting areas that contain healthy, intact ecosystems is an effective


way to preserve overall biodiversity, and some people argue that it is ulti-
mately the only way to preserve many species and ecosystems, particularly
those not well suited for landscapes heavily modified by humans. How-
ever, protecting restored areas can also be valuable and has been shown to
be as good or better for maintaining some ecosystem services (Possingham
et al. 2015). Preserving ecosystems involves:
1. establishing individual protected areas,
2. creating networks of protected areas,
3. managing those areas effectively,
4. implementing conservation measures outside the protected areas, and
5. restoring biological communities in degraded habitats (www.wri.org).
The first three of these topics are discussed here and the final two are
covered in Chapters 9 and 10, respectively.

Establishment and Classification of


Protected Areas
Protected areas can be established in a variety of ways, but the most com-
mon mechanisms are (roughly in decreasing order of significance):
• Government action, usually at a national level, but often on regional
or local levels
Protected Areas 267

• Land purchases and easements by private individuals and conserva-


tion organizations
• Actions of indigenous peoples and traditional societies
• Development of biological field stations (which combine biodiversity
protection and research with conservation education) by universities
and other research organizations

Although legislation and land purchases alone do not ensure habitat protec-
tion, they can lay the groundwork for it. Partnerships among governments
of developing countries, international conservation organizations, multina-
tional banks, research and educational organizations, and governments of
developed countries are another way to bring together funding, training,
and scientific and management expertise to establish new protected areas.
Traditional societies also have established protected areas to maintain
their ways of life or simply to preserve their land (Langton et al. 2014). Many
of these protected areas have been in existence for long periods of time
and are linked to the religious and cultural beliefs of the inhabitants. Na-
tional governments in many countries, including the United States, Canada,
Colombia, Brazil, and Australia, have recognized the rights of traditional
societies to own and manage the land on which they live, hunt, and farm.
However, in some cases, the recognition of land rights results
only after conflict in the courts, in the press, and on the land.
The International Union for Conservation of Nature (IUCN)
The IUCN has developed
has developed a six-category system for classifying protected
areas (Table 8.1). The conservation of nature in protected areas a classification system
is a primary management objective in all six categories (Dudley for protected areas,
2008), with lands in categories I–IV considered strictly protected. ranging from strict nature
However, areas in the fifth and sixth categories are considered reserves to managed-
multiple-use or multi-management protected areas, as they are resource protected areas,
administered, not only to conserve biodiversity, but also to pro- depending on the level
duce natural resources, such as timber and cattle, for human use. of human impact and
These multi-management protected areas can be particularly the needs of society for
significant for several reasons: resources.
• They are often much larger in area than other categories of
protected areas
• They may contain many or even most of their original species
• They often adjoin or surround other protected areas
• They are more likely to benefit local people than strictly protected
areas, and therefore are more likely to earn local support

A review of 171 published reports on 165 protected areas found that all
types (I–VI) were more likely to have a positive impact on the local peo-
ple than a negative impact and that the multi-use protected areas (V and
VI) were the most likely to have positive conservation and socioeconomic
outcomes (Oldekop et al 2015). This is likely because local people benefit
268 Chapter 8

Table 8.1 IUCN Protected Area Designations I–VI


Category Description
Ia. Strict nature reserves Managed mainly for scientific research and monitoring; areas of
land and/or sea possessing some outstanding or representative
ecosystems, geological or physiological features, and/or species
Ib. Wilderness areas Managed mainly for wilderness protection; large areas of
unmodified or slightly modified land and/or sea retaining
their natural character and influence, without permanent or
significant habitation, which are protected and managed so as
to preserve their natural condition
II. National parks Managed mainly for ecosystem protection and recreation; natural
areas of land and/or sea designated to (1) protect the ecological
integrity of one or more ecosystems for present and future
generations; (2) exclude exploitation or occupation inimical
to the purposes of designation of the area; and (3) provide a
foundation for spiritual, scientific, educational, recreational, and
visitor opportunities, all of which must be environmentally and
culturally compatible
III. Natural monuments Managed mainly for conservation of specific natural features;
areas containing one or more specific natural or natural/
cultural features of outstanding or unique value because of
inherent rarity, representative or aesthetic qualities, or cultural
significance
IV. Habitat/species Managed mainly for conservation through management
management areas intervention; areas of land and/or sea subject to active
intervention for management purposes to ensure the
maintenance of habitats and/or to meet the requirements of
specific species
V. Protected landscapes Managed mainly for conservation and recreation; areas of land,
and seascapes with coast and sea as appropriate, where the interaction of
people and nature over time has produced an area of distinct
character with significant aesthetic, ecological, and/or cultural
value, and often with high biological diversity
VI. Managed-resource Managed mainly for the sustainable use of natural ecosystems;
protected areas areas containing predominantly unmodified natural systems,
managed to ensure long-term protection and maintenance of
biological diversity, while also providing a sustainable flow of
natural products and services to meet community needs
Source: After www.iucn.org.

from direct use of the protected area (see Chapter 3) and are not excluded
from its management.
Almost every country currently has one or more protected areas (Jen-
kins and Joppa 2009; www.iucn.org). Countries that protect less than 1% of
their land include Syria, Iraq, Haiti, and Uruguay (UNEP Protected Planet
Report 2014). Although it could be argued that virtually all countries should
have at least one national park, large countries with rich biotas and a vari-
ety of ecosystem types generally benefit from having many protected areas.
Brazil, for example, protects 26% of its land area and has 67 national parks.
Protected Areas 269

(A)

Proportion protected
0% 5 – 10%
< 1% 10 – 17%
1 – 5% > 17%

(B)

Proportion protected
0% 5 – 10%
< 1% > 10
1 – 5%

Figure 8.2 The percentage of the world’s terrestrial and marine ecoregions
that are protected. The darkest color for each map indicates those that have
reached the targets stipulated by the Convention of Biological Diversity. (From
Watson et al. 2014.)

Around 209,000 protected areas in IUCN categories I–VI have been


designated worldwide, in over 190 countries (www.wdpa.org) (Figure
8.2).1 Nevertheless, this impressive number represents only about 15.4% of
Earth’s total land surface, which is about the same area as the land used to
grow all of the world’s crops. Much of this protected land is concentrated
in areas that are at higher elevations, on steeper slopes, and in remote areas
far from roads and cities, where disturbance from humans is often already
minimal (Joppa and Pfaff 2009). For example, the world’s largest park is in
Greenland on inhospitable terrain and covers 970,000 km2, accounting for
about 3% of the global area that is protected. Far more advantageous for
conservation is the establishment of protected areas guided by one of the
approaches for species protection described in Chapter 6. Furthermore,
only about 30% of the protected area is in categories I–IV (strictly protected

1
Uncertainty about the number and size of protected areas stems from the different stan-
dards used throughout the world, the degree of protection actually given to a particular
designated area, and variations in when the data were gathered.
270 Chapter 8

in scientific reserves and national parks), with the greatest growth in pro-
tected areas that share priorities for human use (Figure 8.3).
The measurements of protected areas in individual countries and on
continents are only approximate because sometimes the laws protecting
national parks and wildlife sanctuaries are not strictly enforced. At the
same time, there are sections of managed areas that, while not legally pro-
tected, are carefully protected in practice. Examples include the designated
wilderness areas within US national forests that forbid logging, grazing,
mountain bikes, and motorized vehicles.
The proportion of land that is strictly protected varies dramatically
among countries: high proportions of land are protected in Germany (42%),
Austria (23%), and the United Kingdom (26%), and surprisingly low propor-
tions are protected in Bosnia and Herzegovina (1%), Ireland (2%), Hungary
(5%), and Denmark (5%). Moreover, even if a country has numerous protected
areas, certain unique habitats that also have high economic value may re-
main unprotected. In addition, protected areas may be reduced in size by the
government, opened up for exploitation, or even have their protected status
removed (known as degazettement), particularly if they are found to contain
valuable natural resources (Mascia and Pallier 2011; Mascia et al. 2014).
The limited extent of protected areas highlights the biological sig-
nificance of the more than 23% of the world’s land that is managed for
sustainable resource production, such as production forests, watersheds

Ia: Strict nature reserve


25 Ib: Wilderness area
II: National park
Protected areas (million km2)

20 III: Natural monument


IV: Habitat species management
V: Protected landscape/seascape
15
VI: Protected areas with sustainable
use of natural resources
10

0
1950 1960 1970 1980 1990 2000 2010 2014
Year

Figure 8.3 The increase in amount of area in both MPAs and terrestrial
protected areas assigned to each of the six IUCN categories over time. Colored
areas are those for which data has been provided to the World Database on
Protected Areas (WDPA) and represents only 64% of total global protected area;
those for which categories have not been assigned were not included. (After
UNEP-WCMC Report 2014.)
Protected Areas 271

around reservoirs, and grazing lands, which is described in greater detail


in Chapter 9.

Marine protected areas (MPAs)


Marine conservation has lagged behind terrestrial conservation efforts;
however, rapid progress is now being made in MPAs (Gaines et al. 2010).
Only about 10.9% of the world’s territorial seas near coastlines and about
3.4% of the total marine environment are included in protected areas
(Figure 8.4). The international goal of protecting 10% of the entire marine

Key to marine protected areas


No-take marine reserve; officially designated and implemented
No-take marine reserve; officially designated, not implemented
Multi-zoned (with some no-take) marine proteced area; officially designated and implemented
Multi-zoned (with some no-take) marine proteced area; officially designated, not implemented
Other MPAs

Figure 8.4 The world’s terrestrial protected areas (shown in light brown) and
Marine Protected Areas (see key). Although many small protected areas do not
show up at this scale, large areas in IUCN categories I–VI are indicated, as well as
many areas that are protected in some manner (e.g., privately) but do not have an
official designation at the present time. Note the large protected areas in Green-
land, western Europe, eastern and western North America, the Amazon Basin,
northeastern Australia, and western China. Many large new MPAs have recently
been designated (numbered circles), including (1) the expansion of the Pacific
Remote Islands National Monument in 2014, (2) new areas around Pitcairn Islands
in 2015, and (3) Gabon in 2014. (From MPAtlas.org. [Current 02/2016].)
272 Chapter 8

environment by the year 2020 was motivated in part to manage declining


commercial fishing stocks (Rife et al. 2013); however, even stronger mea-
sures may be required to conserve the full range of coastal and marine
biodiversity (Spalding et al. 2008; www.iucn.org). Some of the current pro-
posals for new MPAs are receiving pushback from groups that fear it will
harm the fishing industry. One such example is protests of the proposal
to introduce 30 new MPAs off Scotland in 2016, even though MPAs can
benefit fisherman by providing opportunities for fish to reproduce and
grow larger.
Over 5000 marine and coastal protected areas have been established
worldwide, but most are small. A number of countries, though, are now
creating very large marine protected areas (VLMPAs) covering hundreds
of thousands of square kilometers of marine ecosystems. One of the larg-
est single reserves was established in 2015 around the Pitcairn Islands
in the South Pacific Ocean; 830,000 km 2 was set aside as a no-take zone,
with only traditional fishing by locals permitted. This reserve is dwarfed,
however, by the Pacific Remote Islands Monument, a series of areas that
was expanded to eight times their original area by President Obama in
2014. It now covers a total of 2 million km 2.
Marine protected areas seek to preserve the nursery grounds of com-
mercial species and to maintain water quality and both the physical and
biological features of ecosystems (Gaines et al. 2010). In the process, high-
quality protected areas can also maintain recreational activities such as
swimming and diving and the economic benefits associated with tourism.
Unfortunately, many MPAs exist only on maps and receive insufficient
regulation, funding, and enforcement to prevent overharvesting and pol-
lution. Even in the well-funded Phoenix Islands Protected Area, located in
the central Pacific, fishing is only banned in 3% of the total area.

The effectiveness of protected areas


Studies show that protected areas generally are effective at protecting eco-
systems, particularly when regulations are enforced and when areas are
geographically isolated (Geldmann et al. 2013). The value of protected areas
in maintaining biodiversity is abundantly clear in many tropical countries
(Nolte et al. 2013) (see Figure 4.6). Inside a given park’s boundaries you may
see numerous trees and abundant animal life, while outside the park you
may very well see cleared land and see and hear few animals (Figure 8.5).
Many of these protected areas also continue to provide income and services

Figure 8.5 National parks and other protected areas are able to prevent
damage to the natural forests in (A) the Atlantic coast forests of Brazil and (B)
West Africa. In the Atlantic coast forest, there is a sharp boundary with intact
forest inside the protected areas and around 50% intact forest outside the pro-
tected area. For protected areas in West Africa, there is considerable forest deg-
radation within 16 km of the park boundary, particularly for IUCN categories V
and VI, which are mainly forest reserves. See Table 8.1 for a description of IUCN
categories. (After Joppa et al. 2008.)
Protected Areas 273

to poor people living nearby (Andam et al. 2010). At the same time, these
areas must be monitored and managed to protect them from overharvest-
ing by both legal and illegal subsistence use and commercial production. In
some cases, due to management problems and conflicts with local inhabit-

(A) Brazil
100

80
Natural vegetation (%)

Inside the park


60

40

Outside the park

20 IUCN I–II
IUCN III, IV
IUCN V, VI

0
–24 –16 –8 0 8 16 24
Distance from park boundary (km)
(B) West Africa
100

80

Outside the park


Natural vegetation (%)

60

40
Inside the park

20

0
–24 –16 –8 0 8 16 24
Distance from park boundary (km)
274 Chapter 8

ants, national parks have become more degraded than neighboring areas
(Wright et al. 2007). Consequently, the effective management of protected
areas must consider activities that occur in neighboring areas. There are
also cases where the protected status is violated by the very governments
charged with enforcing it, such as the extensive logging in the Sochi Na-
tional Park in the Western Caucasus to enable construction for the Olympic
Games (Bragina et al. 2015).
If national parks can be established and maintained where
concentrations of species occur, high numbers and percent-
ages of species can be preserved (Joppa et al. 2013). A system
Although the number of protected areas can include a high percentage of a country’s
of species living within species if it includes representatives of all major ecosystems.
a protected area is an For example, in Britain, 88% of plant species occur in the pro-
important indicator of the tected area system, of which 26% are found exclusively within
area’s potential to protect protected areas (Jackson et al. 2009). Likewise, China’s nature
biodiversity, protected reserves cover 15% of the total area and include 81% of the coun-
areas need to maintain try’s vegetation types (Wu et al. 2011). However, it is important
healthy ecosystems and to recognize that the long-term survival of many species in pro-
viable populations of tected areas, and even of the ecosystems themselves, remains
important species. in doubt because populations of many species and the area of
the ecosystems may be so reduced in size that their eventual
fate is extinction.

Measuring effectiveness: Gap analysis


Gap analysis compares biodiversity priorities with existing and proposed
protected areas (see, for example, Simaika et al. 2013), but it can consider
several factors. In the past, conservationists used informal means to ensure
that the high-priority areas were protected, for example, by establishing
national parks in different regions with distinctive ecosystems and ecologi-
cal features (Shafer 1999). Now, however, conservationists are using more
systematic planning processes involving gap analysis (Tognelli et al. 2009).
Gap analysis generally consists of the following steps:
1. Data are compiled describing the presence and distribution of species,
ecosystems, and physical features of the region, which are sometimes
referred to as conservation units. Information on human densities and
economic factors can also be included.
2. Conservation and social goals are identified, such as the amount of area
to be protected for each ecosystem, the number of individuals of rare
species to be protected, or the desired balance between wilderness and
mixed resource management.
3. Existing conservation areas are reviewed to determine what is protected
already and what is not (known as “identifying gaps in coverage”).
4. Additional areas are identified to help meet the conservation goals
(“filling the gaps”).
Protected Areas 275

5. These additional areas are reviewed in more detail and, if appropriate


and practical, protected in some way (often by being directly purchased
or designated as national parks). Management plans are then developed
and implemented.
6. The new protected areas are monitored to determine whether they are
meeting their stated goals. If not, the management plan can be changed
or, possibly, additional areas can be acquired to meet the goals.

Gap analysis can also be used to identify holes in conservation at in-


ternational scales. For example, a group of researchers from 27 institu-
tions and organizations did an analysis of all existing protected areas
and 25,380 species across the globe (Butchart et al. 2015). They calculated
that protected areas currently only include 77–78% of important sites for
biodiversity and there is insufficient coverage for 57% of the species they
evaluated (Figure 8.6). They determined that in order to reach the Aichi
Biodiversity targets of >17% and >10% marine systems protected, an ad-
ditional 3.3 million km2 outside existing protected areas were needed, plus
14.8 million km 2 to cover all threatened species. To cover these plus all
documented important sites for biodiversity and individual country and
biome targets, the researchers estimated that the amount of area protected
needed to be roughly doubled (see Figure 11.7).

Complete protection Unprotected


100 Partial protection
Proportion in each category (%)

80

60

40

20

0
A
M

M ny

C she

Se
Bi

M
Lo nd
m

ar s

or
am

ar fi
rd

bo

fi

ag
an
a
ph

bs cra
til

al
in sh
s

r
m

gr
te yfi
ag

s
ib

as
e
al

rs s

ov
i

se
in
an
s

es

s
o
s

us

Threatened species within a group

Figure 8.6 A gap analysis for various groups of animals and plants shows
the proportion of threatened species that receive any sort of protection. Depend-
ing on the group, 10–30% of threatened species lack protection anywhere in
their range; most threatened species have at least partial protection. Relatively
few threatened species are completely protected. (After Butchart et al. 2015.)
276 Chapter 8

Geographic information systems (GIS) are a vital tool in gap


GIS is an effective tool analysis; they facilitate the integration of the wealth of data on
for gap analysis, which the natural environment with information on species distribu-
uses a wide variety of tions (Murray-Smith et al. 2009). The basic GIS approach involves
information to pinpoint storing, displaying, and manipulating many types of spatial
critical areas and species data involving factors such as vegetation types, climate, soils,
that are priorities for topography, geology, hydrology, species distributions, human
protection. settlements, and resource use (Figure 8.7). This approach can
point out correlations among the abiotic, biotic, and human ele-

Vegetation
types

A
Distributions of
endangered
animal
species B

Preserves

A
Overlapped maps
show gaps in
protection B

Figure 8.7 Geographic information systems (GIS) provide a method for inte-
grating a wide variety of data for analysis and display on maps. In this example,
vegetation types, distributions of endangered animal species, and preserved
areas are overlapped to highlight areas that need additional protection. The
overlapped maps show that the distribution of Species A is predominantly in a
preserve, Species B is only protected to a limited extent, and Species C is found
entirely outside the preserves. Establishing a new protected area to include the
range of Species C would be the highest priority. (After Scott et al. 1991.)
Protected Areas 277

ments of the landscape; help plan parks that include a diversity of biological
communities; and even suggest sites that are likely to support rare and
protected species. Aerial photographs and satellite imagery are additional
sources of data for GIS analysis, and they can highlight patterns of veg-
etation structure and distribution over local and regional scales. These
images can dramatically illustrate when current government policies are
not working and need to be changed.

Designing Protected Areas


The size and placement of protected areas throughout the world are often
determined by the distribution of people, potential land values, the political
efforts of conservation-minded citizens, and historical factors (Armsworth
et al. 2006; Mills et al. 2014). Ideally, the selection of new protected areas
will follow conservation priorities, such as to protect a particular species,
ecosystem, or biodiversity hotspot (see Chapter 6); however, there are many
other possible scenarios. In urban and suburban areas, the fund-raising
ability of private conservation groups, wealthy individuals, and govern-
ment departments is often the most important factor in determining what
land is acquired (Lerner et al. 2007). In other cases, certain parcels of land
may be purchased to protect a critical water supply or a charismatic species,
but yet others may be acquired simply because they adjoin the property of
an influential citizen. Moreover, sometimes lands are set aside for conser-
vation protection because they have no immediate commercial value—they
are “the lands that nobody wants” (Scott et al. 2001).
Issues of reserve design have proved to be of great interest to gov-
ernments, corporations, and private landowners, who are being urged,
or even mandated, to manage their properties for both the commercial
production of natural resources and the protection of biodiversity. How-
ever, conservation biologists must be cautious about using simplistic,
overly general guidelines for designing protected areas, because every
conservation situation requires individual consideration. Everyone ben-
efits from increased communication between the academic scientists,
who are developing theories of nature reserve design, and the manag-
ers, planners, and policymakers, who are actually creating new reserves
(Braunisch et al. 2012).
Conservation biologists often start by considering the four Rs:
1. Representation. Protected areas should contain as many features of bio-
diversity (species, populations, habitats, etc.) as possible.
2. Resiliency. Protected areas must be sufficiently large to maintain all as-
pects of biodiversity in a healthy condition for the foreseeable future,
including as climate conditions change.
3. Redundancy. Protected areas must include enough examples of each
aspect of biodiversity to ensure its long-term existence in the face of
future uncertainties.
278 Chapter 8

4. Reality. There must be sufficient funds and political will, not only to
acquire and protect lands, but also to regulate and manage the pro-
tected areas.
The following, more specific, questions about reserve establishment and
design are also useful for discussing how best to construct and link pro-
tected areas:
• Given a particular amount of funding to spend on a protected area or
network of areas, what is the most effective way to spend it?
• How large must a nature reserve be to effectively protect biodiversity?
• Is it better to have a single large protected area or multiple smaller
reserves?
• When a network of protected areas is created, should the areas be far
apart or close together, and should they be isolated from one another
or connected by corridors?
• How many individuals of an endangered species must be included in
a protected area to prevent the local extinction of a species?
• What is the most cost-effective way to design a protected area to
achieve its conservation goals?
• What is the best shape for a nature reserve?

Conservationists explore many of these issues using the


Principles of design island biogeography model of MacArthur and Wilson (1967),
have been developed to which describes the relationship between the size of an area and
guide land managers the number of species it supports (see Chapter 5). When applied
in establishing and to protected areas, the model frequently assumes that parks are
maintaining networks of habitat islands —intact habitat surrounded by an unprotected
protected areas. matrix of inhospitable terrain. In fact, many species are capable
of living in and dispersing through the surrounding habitat
matrix (see Chapter 9). This may explain the finding that species
diversity on oceanic islands increases more rapidly with size
than do habitat islands, and that among habitat islands, this relationship
is stronger for urban islands than for forest islands (Matthews et al. 2015).
Researchers working with island biogeography models and data from pro-
tected areas have proposed various principles of reserve design, which are
still being debated (Figure 8.8).

Figure 8.8 Principles of reserve design that are based in part on theories
of island biogeography. Imagine that the reserves are “islands” of the original
biological community surrounded by land that has been made uninhabitable for
the original species by human activities such as farming, ranching, or industrial
development. The practical application of these principles is still being studied
and debated, but in general, the designs shown on the right are considered pref-
erable to those shown on the left. (After Shafer 1997.)
Protected Areas 279

Worse Better

River
(A) Ecosystem Ecosystem
partially completely
protected protected

(B) Smaller Larger


reserve reserve

(C) Fragmented Unfragmented


reserve reserve

(D) Fewer More


reserves reserves

(E) Isolated Corridors


reserves maintained

(F) Isolated “Stepping-stones“


reserves facilitate movement

(G) Uniform Diverse habitats


habitat (e.g., mountains,
protected lakes, forests)
protected
300 ha reserve 100 ha
core
(H) Irregular Reserve shape
shape closer to round
(fewer edge effects)
300 ha reserve

(I) Only large Mix of large and


reserves small reserves

Reserves managed
(J) Reserves regionally
managed
individually

Stop
(K) Humans Human integration;
excluded buffer zones
280 Chapter 8

Also, aspects of protected area models differ depending on whether


the targets are terrestrial or marine species; vertebrates, higher plants, or
invertebrates; or whole ecosystems. Many aspects of marine species’ life
cycles and dispersal mechanisms are largely unknown (Burgess et al. 2014;
Schofield et al. 2013). MPAs and complementary conservation efforts out-
side protected areas must be designed to accommodate the high mobility
of some species and limited dispersal of other target species, thus ensuring
that corridors and genetic populations are preserved.

Protected area size and characteristics


An early debate in conservation biology centered on one ques-
tion: Is species richness maximized in one large nature reserve
Large reserves are or in several smaller ones of an equal total area (McCarthy et al.
generally better able 2006; Soulé and Simberloff 1986)? This debate is known in the
to maintain many literature as the SLOSS debate (with “SLOSS” standing for single
species because such large or several small). Is it better, for example, to set aside one
reserves support larger reserve of 10,000 ha or four reserves of 2500 ha each? Proponents
population sizes and
argue that only large reserves can maintain sufficient numbers
of large, wide-ranging, low-density species (such as large carni-
a greater variety of
vores) over the long term. Large reserves also minimize the ratio
habitats. However, small
of edge habitat to total habitat, encompass more species, and
reserves are important
sometimes have greater habitat diversity than small reserves.
in protecting particular
The advantage of large parks was effectively demonstrated
species and ecosystems. by an analysis of 299 mammal populations in 14 national parks
in western North America (Figure 8.9) (Newmark 1995). Ex-
tinction rates were very low or zero in parks with areas over

0.004
Lassen Volcanic

Figure 8.9 Mammals have higher Bryce


extinction rates in smaller parks than 0.003
in larger parks. If these same trends
continue for the next 100 years, small Crater Lake
Extinction rate

parks such as Bryce are predicted to Mount Rainier


Manning
lose more than 30% of their mammal 0.002 Provincial
species, whereas large parks such Zion
as Yosemite will only lose 5%, and
Grand Canyon
Kootenay-Banff-Jasper-Yoho is not
expected to lose any species. Each dot 0.001 Yosemite
Sequoia-Kings Canyon
represents the actual extinction rate Grand Teton
of mammal populations (expressed Glacier-Waterton Lake Kootenay-Banff-
as the proportion of species that have Rocky Mt. Olympic Jasper-Yoho
gone extinct per year) for a particular 0.000
US national park, Canadian national 100 1,000 10,000 100,000
park, or two or more adjacent parks. Small Large
(After Newmark 1995.) parks Park area (km2) parks
Protected Areas 281

1000 km 2 but much higher in parks that are smaller than 1000 km 2. In
addition to the ability to support more species because of their larger size,
larger parks tend to be surrounded by lower densities of people compared
to small reserves, potentially increasing their connectivity with other eco-
systems and populations (Wiersma et al. 2004).
On the other hand, once a park reaches a certain size, the number of
new species added with each increase in area starts to decline. At that
point, creating a second large park or another park some distance away
may be a more effective strategy for preserving additional species than
simply adding area to the existing park.
The research on extinction rates of populations in large parks has three
practical implications:
1. When a new park is being established, it should generally be made
as large as possible (within the context of an overarching strategy for
optimizing protected areas) in order to preserve as many species as
possible, contain large populations of each species, and provide a di-
versity of habitats and natural resources. Keystone resources should
be included, in addition to habitat features that promote biodiversity,
such as elevational gradients.
2. When possible, land adjacent to protected areas should be acquired to
reduce external threats to existing parks and maintain buffer zones. For
example, terrestrial habitats adjacent to wetlands are often needed by
semiaquatic species such as snakes and turtles. Moreover, protecting
natural ecological units, such as entire watersheds or mountains, is
often the best means to reduce external threats.
3. The effects of climate change, invasive species, and other threats are
altering ecosystems within existing protected areas. These changes can
reduce the area of habitat available for a species and lead to declines in
population size and increased probability of extinction. These changes
emphasize the need for preserving corridors or otherwise connecting
protected areas to facilitate the dispersal of species among them.

Although research suggests that large parks may be best able to achieve
conservation goals in many situations, there are some instances where
several small parks or a mixture of large and small ones is better for con-
servation. For example, a study of wetlands in New Zealand determined
that certain rare plant species were not necessarily present in the large
protected areas and that several small reserves were more effective for
their protection (Richardson et al. 2015).
Small reserves, even those less than a hectare in size, may effectively
protect isolated populations of rare species, particularly if they contain a
unique habitat type (Jarošík et al. 2011). Also, regional biodiversity may
depend on small natural features (SNFs) such as temporary pools, caves,
single trees, or rock outcrops (Hunter et al. 2015). These SNFs may contain
keystone resources (discussed later) or ecological processes and as such
282 Chapter 8

have a disproportionately large conservation value, a phenomenon that


has been called the Frodo effect.2 For example, ephemeral desert springs
should be prioritized for inclusion in protected areas because so many
organisms are dependent on their cycles of wet and dry periods (Acuña
and Ruhí 2016). Thus, several small protected areas that include critical
SNFs will conserve more species than a single large one that does not
contain those SNFs.
Even if a large reserve would protect the most species in a given system,
often officials have no choice but to accept the challenge of managing spe-
cies and biological communities in small reserves. The 10,000 protected
areas in Britain, for instance, have an average area of only 3 km 2 (Figure
8.10) (Jackson et al. 2009). Numerous countries have many more small
protected areas (less than 100 ha) than medium and large ones, yet the com-
bined area of these small reserves is
only a tiny percentage of the total area
under protection. This is particularly
European areas
true in places, such as Europe, China,
International wetlands
and Southeast Asia, that have been in-
National areas
tensively cultivated for centuries.
Singapore provides an excellent
example of how small reserves pro-
vide long-term protection for numer-
ous species. A group of reserves that
have been isolated since 1860 currently
protect 5% the original habitat on Sin-
gapore, yet they still contain around
half of the country’s original flora and
fauna, including 350 of the original
bird species and 26 of the mammal
species (Corlett 2013). In addition,

Figure 8.10 The geographic loca-


tions of protected areas in Britain man-
aged for biodiversity conservation. Note
their large number, varied sizes and
shapes, and scattered distribution. At
the scale of this map, most of the small
protected areas cannot be seen. Many of
the protected areas are covered by two
or more designations. There are other
areas managed for other purposes, which
are not shown on this map. (Courtesy of
Sarah Little.)

2
The Frodo effect borrows its name from the J. R. R. Tolkien character in the Lord of the
Rings, who is small in stature yet responsible for the future of everyone else.
Protected Areas 283

small reserves located near populated areas make excellent conservation


education and nature study centers that further the long-range goals of
conservation biology by developing public awareness of important issues.
By 2030, over 60% of the world’s population will live in urban areas; thus,
there is a need to develop such reserves for public use and education.
It is generally agreed that protected areas should be designed to mini-
mize edge effects. Conservation areas that are rounded in shape minimize
the edge-to-area ratio, and the center is farther from the edge than in other
park shapes. Long, linear parks have the most edge—all points in the park
are close to the edge (Yamaura et al. 2008). Consequently, for parks with
four straight sides, a square park is a better design than an elongated rect-
angle of the same area. However, most parks have irregular shapes because
land acquisition is typically a matter of opportunity rather than design.
As discussed in Chapter 4, the internal fragmentation of protected
areas by roads, fences, farming, logging, and other human activities
should be avoided as much as possible because fragmentation creates
barriers to dispersal and often divides a large population into two or
more smaller populations, each of which is more vulnerable to extinc-
tion than the large population. Fragmentation also provides entry points
for invasive species, which may harm native species, and creates more
undesirable edge effects.

Networks of Protected Areas


To overcome some of the effects of fragmentation and limits on the size of
protected areas, conservationists have developed strategies to aggregate
small and large protected areas into larger conservation networks (Bode
et al. 2011; Van Teeffelen et al. 2012). Networks are also important to ac-
commodate metapopulation dynamics (see Chapter 6).
Cooperation among public and private landowners is important for
creating these networks, particularly in developed metropolitan areas
and other areas where there are many small, isolated parks controlled
by different government agencies and private organizations. An excellent
example of cooperation intended to network urban protected areas is the
Chicago Wilderness project, which consists of more than 300 organizations
collaborating to preserve around 227,000 ha of tallgrass prairies, wood-
lands, rivers, streams, and other wetlands in metropolitan Chicago (www.
chicagowilderness.org).
Often a protected area is embedded in a larger matrix of habitat man-
aged for human uses such as timber forest, grazing land, and farmland
(Figure 8.11). If the protection of biodiversity becomes a secondary prior-
ity of these areas and is included in their management plans, then these
mixed-use areas can serve as buffers, corridors, or other key components of
protected area networks (Hansen et al. 2011). Habitat managed for resource
extraction can sometimes also be designated as important secondary sites
for wildlife and as dispersal corridors between isolated protected areas.
284 Chapter 8

Figure 8.11 Tina Buijs, a


park guard supervisor and opera-
tions manager with The Nature
Conservancy (TNC), talks with
Juan Antillanca, a farmer be-
longing to the Huiro indigenous
community that borders TNC’s
Valdivian Coastal Reserve in
Chile. The reserve is a 61,000 ha
site comprising temperate rain
forest and 36 km of Pacific coast-
line. Keeping in close contact with
their neighbors helps TNC officials
realize their conservation goals.
(Photograph © Mark Godfrey/The
Nature Conservancy.)

Using a network approach, groups of rare species can be managed as large


metapopulations to facilitate gene flow and migration among populations
(Andrello et al. 2015).

Habitat corridors
Growing numbers of conservationists argue that connectivity is important,
and they are taking steps to link isolated protected areas into large systems
through the use of habitat corridors—strips of protected land running be-
tween the reserves (Beier 2011; Magrach et al. 2012). Such habitat corridors,
also known as conservation corridors or movement corridors, can allow
plants and animals to disperse from one reserve to another, facilitating
gene flow and the colonization of suitable sites.
Corridors are clearly needed to preserve animals that must migrate
seasonally among different habitats to obtain food and water, such as the
large grazing mammals of the African savanna. If these animals were
confined to a single reserve by fences, farms, and other anthropogenic
factors, they might starve (Wilcove and Wikelski 2008). The width required
for effective corridors varies depending on the species, length of the cor-
ridor, and other factors.
In many areas, roads are a primary obstruction to the cre-
ation of habitat corridors. In these cases culverts, tunnels, and
overpasses can create passages under and over roads and rail-
Establishing habitat
ways that allow reptiles, amphibians, and mammals to travel
corridors can potentially
between habitat fragments or protected areas (Soanes et al. 2013).
transform a set of isolated
An added benefit of these passageways is that they reduce col-
protected areas into a lisions between animals and vehicles, which saves lives and
linked network with money. For example, in Canada’s Banff National Park, road col-
populations interacting lisions involving deer, elk, and other large mammals declined
as a metapopulation. by 96% after fences, overpasses, and underpasses were installed
along a major highway (Figure 8.12A) (Ford et al. 2009).
Protected Areas 285

Corridors that facilitate natural patterns of migration will probably


be the most successful at protecting species. For example, large grazing
animals often migrate in regular patterns across a rangeland in search of
water and the best vegetation. In seasonally dry savanna habitats, animals
often migrate along the riparian forests that grow along streams and rivers.
In mountainous areas, many bird and mammal species regularly migrate
to higher elevations during the warmer months of the year. For example,
a corridor was established in Costa Rica to link two wildlife reserves, the
Braulio Carrillo National Park and La Selva Biological Station, to protect
migrating birds. A 7700 ha corridor of forest several kilometers wide and
18 km long, known as the Zona Protectora Las Tablas, was set aside to
provide an elevational link that allows at least 75 species of birds to migrate
between the two large conservation areas (Bennett 1999).
Some conservation biologists have started to plan habitat corridors on a
truly huge scale. Wildlands Network has a detailed plan, called the Spine
of the Continent Initiative, that would link all large protected areas in the
western United States and Canada by habitat corridors, creating a system
that would allow large and currently declining mammals to coexist with
human society.

(A) Figure 8.12 (A) An overpass above a


fenced-off divided highway allows animals to
migrate safely between two forested areas.
(B) Individuals of a species naturally disperse
between two large protected areas (areas 1 and
2, left-hand panel) by using smaller protected
areas as stepping-stones. The right-hand panel
shows that habitat destruction and a large
edge effect zone caused by a new road have
blocked a migration route. To offset the effects
of the road, compensation sites (orange) have
been added to the system of protected areas,
and an overpass has been built over the high-
way to allow dispersal. (A, © Paul Zizka/Getty
Images; B, after Cuperus et al. 1999.)

(B) Before fragmentation After fragmentation

1 1

Highway
Effect zone
Overpass
2 Nature area
2
Compensation site
Dispersal routes
286 Chapter 8

As the global climate changes, many species are moving to higher


elevations and higher latitudes. Creating corridors to protect expected
routes—such as north–south river valleys, ridges, and coastlines—would
be a useful precaution. Extending existing protected areas in the direc-
tion of anticipated species movements could help to maintain long-term
populations (Nuñez et al. 2013). Corridors that cross gradients of elevation,
rainfall, and soils could also allow the local migration of species to more
favorable sites.
Although the idea of corridors is intuitively appealing, there are some
possible drawbacks (Ogden 2015; Orrock and Damschen 2005; Simberloff
et al. 1992). In particular, corridors may facilitate the movement of pest
species and disease; a single infestation could quickly spread to all of the
connected nature reserves and cause the extinction of all populations of
a rare species. Also, animals dispersing along corridors may be exposed
to greater risks of predation because human hunters as well as animal
predators tend to concentrate on routes used by wildlife. Finally, there is a
risk that the corridors, which in some cases can be expensive and difficult
to construct, will not be used by the intended species.
Some studies published to date support the conservation value of
corridors, while other studies show no effect (Gilbert-Norton et al. 2010;
Pardini et al. 2005). In general, maintaining existing corridors is probably
worthwhile because many are located along watercourses that may be bio-
logically important habitats themselves. New protected corridors are most
obviously needed along known migration routes (Newmark 2008). There
would also be value in leaving small clumps of original habitat between
large conservation areas to facilitate movement in a stepping-stone pattern
(Figure 8.12B), such as the protected areas that birds use as stopping
points along the flyways of their annual migration routes.

Landscape Ecology and Park Design


The interaction of human land-use patterns, conservation theory, and park
design is evident in the discipline of landscape ecology, which investi-
gates patterns within the mosaic of the physical environment, ecological
communities, ecosystem processes, and human–ecosystem interactions
on local and regional scales (Schwenk and Donovan 2011; Wu and Hobbs
2009) (Table 8.2). It has focused on terrestrial systems, but the
same approach can apply to marine and other aquatic systems
(Jelinski 2015).
In some cases, long-term Landscape ecology has been most intensively studied in
traditional human use Europe and Asia, where long-term practices of traditional ag-
has created landscape riculture and forest management have shaped landscape pat-
patterns that preserve terns. In the European countryside, cultivated fields, pastures,
and even increase bio- woodlots, and hedges alternate to create a mosaic that affects the
diversity. distribution of wild species. Likewise, in the traditional Japanese
landscape known as satoyama, flooded rice fields, hay fields, vil-
Protected Areas 287

Table 8.2 A Proposed Framework for Landscape Design That


Promotes Biodiversity Conservation
1. Distinguish and delineate different patches of land covers (e.g., rice fields, hay
fields, forest, villages) in the selected landscape.
2. Categorize patches as unaltered (lower human use) vs. altered (higher human
use) land covers.
3. Identify the constraints on land use planning (e.g., economic, social, political).
4. Given these constraints, create a landscape plan that maximizes the total
amount and diversity of unaltered land cover, especially near water (e.g., by
restoring some abandoned hay fields to forest);
5. …minimize human disturbance within altered land cover, especially near water
(e.g., by removing roads through forest);
6. …and aggregate altered land covers associated with high-intensity land uses,
especially away from water (e.g. by planning future development to be near
existing villages, rather than in currently unaltered areas).
Source: After Gagné, et al. 2015.

lages, and forests provide a rich diversity of habitat for wetland species, such
as dragonflies, amphibians, and waterfowl (Kadoya et al. 2009) (Figure
8.13). These heterogeneous landscapes, which include a mix of human-
created and natural features, are critical for the survival of some species. In
many areas, traditional patterns of farming, grazing, and forestry are being
abandoned. In some places, rural people have left the land completely and
migrated to urban areas or their farming practices have become more inten-
sive, involving more machinery and the application of fertilizer. To protect
species and ecosystems in such situations, the design and management
of protected areas frequently include strategies to
maintain the traditional landscapes, in some cases
by subsidizing traditional practices or having vol-
unteers manage the land.
To increase the number and diversity of ani-
mals, wildlife managers sometimes create the
greatest amount of landscape variation possible
within the confines of some protected areas, par-

Figure 8.13 Schematic of a traditional rural


landscape near Tokyo, Japan, with an alternating pat-
tern of villages (black); secondary forest (dark green);
paddies, or wet rice fields (light green); and hay fields
(beige). Such landscapes were common in the past
but are now becoming rare because of the increasing
mechanization of Japanese agriculture, the move-
ment of the population away from farms, and the
urbanization of the Tokyo area. The area cover is ap-
proximately 4 km × 4 km. (After Yamaoko et al. 1977.) 1 km
288 Chapter 8

ticularly refuges or other areas managed primarily for hunting and fish-
ing. Fields and meadows are created and maintained, small thickets are
encouraged, groups of fruit trees and crops are planted, patches of forests
are periodically cut, small ponds and dams are developed, and numer-
ous trails and dirt roads meander across and along all the patches. Such
landscaping is often appealing to the public, who are the main visitors and
financial contributors to the park. However, the species in these landscapes
are likely to be principally common species that depend on human dis-
turbance—and in some cases, invasive species. To remedy this localized
approach, large animals, such as bears, mountain lions, and tigers, gener-
ally are best managed on the level of a regional landscape, in which the
sizes of the landscape units more closely correlate to the natural population
sizes and migration patterns of the species (Wikramanayake et al. 2011).

Managing Protected Areas


Some people believe that “nature knows best” and that humans do not
need to actively manage biodiversity—that once protected areas are legally
established, the work of conservation is largely complete. The reality, how-
ever, is often very different. Management is required because humans have
already modified many local environments so much that the remaining
species and ecosystems need human monitoring and intervention in order
to survive. The world is littered with “paper parks” created by government
decree but left to flounder without any management (Joppa et al. 2008).
These protected areas have gradually—or sometimes rapidly—lost species
as their habitat quality has degraded. In some countries, people readily
farm, log, mine, hunt, and fish in protected areas because they feel that
government land is owned by “everyone,” so anybody can take whatever
they want and nobody is willing to intervene.
Many parks are actively managed according to carefully prepared man-
agement plans designed to prevent deterioration. Important considerations
in management plans include protecting biodiversity, maintaining ecosys-
tem services and health, maintaining historical landscapes, and providing
resources and experiences of value to local inhabitants and visitors (Hobbs
et al. 2010). Part of the management plan involves making the public aware
of which activities are encouraged (for example, wildlife photography)
and which are prohibited (for example, hunting), and then enforcing the
rules. The public can even be encouraged to engage in behaviors that help
mitigate the negative impacts of ecotourism (Figure 8.14).
In many European, Asian, and African countries with well-established
traditions of cultivation, ranching, and grazing, hundreds (and even thou-
sands) of years of human activity have shaped habitats such as woodlands,
meadows, and hedges. These habitats support high species diversity as a
result of traditional land-management practices, which must be maintained
if the species are to persist (Jacquemyn et al. 2011). Similarly, grasslands
Protected Areas 289

Figure 8.14 Management of protected areas may involve advising visitors


how to avoid harming wildlife, as does this sign in the parking lot of Table Moun-
tain National Park, on the Cape of Good Hope in South Africa. (Photographs by
Anna Sher.)

that have been grazed in the past by large wild animals or domesticated
animals, such as cattle, still need to be grazed. If protected areas that in-
clude these types of habitats are not managed, they will undergo ecological
succession (a predictable, gradual and progressive change in species over
time) and many of their characteristic species will disappear as shrubs
trees become dominant (Figure 8.15).
It is important, though, to be cautious in taking management actions.
In some cases, often because of a lack of complete understanding of an
ecosystem or conflicting management objectives, management practices
may be ineffective or even detrimental. Some protected areas are managed
to promote the abundance of a game species, such as deer, for hunting.
Management has frequently involved eliminating top predators, such as
wolves and cougars. However, without predators to control them, game
populations (and, incidentally, rodents that feed on seeds and can spread
disease) sometimes increase far beyond expectations, resulting in overgraz-
ing, habitat degradation, and a collapse of animal and plant communities.
Overenthusiastic park managers who remove hollow or dead trees, rot-
ting logs, and underbrush to “improve” a park’s appearance may unwittingly
remove a critical keystone resource needed by certain animal species for
nesting and overwintering, by rare plants for seed germination, and by all
species as an integral part of nutrient cycling (Keeton et al. 2007). In these
instances, a “clean” park can become a biologically sterile park. Likewise,
in many parks, fire is part of the natural ecology of the area (Nimmo et al.
2013). Attempts to suppress fire completely are expensive and waste scarce
management resources. Suppressing the normal fire cycle may eventually
lead to the loss of fire-dependent species and to massive, uncontrollable fires.
290 Chapter 8

Number of plant species/10 m2


40 Bison
Cattle
Ungrazed

30

20

19

19

19

19

19

20

20

20

20

20
95

96

97

98

99

00

01

02

03

04
Year

Figure 8.15 Large herbivores originally grazed


the tallgrass prairies of the midwestern United States.
The loss of these herbivores has altered the ecology of
this ecosystem, with a resulting loss of plant species.
Management involving grazing by cattle and bison
resulted in a gradual increase in plant species in prairie
research plots over a 10-year period, compared with
ungrazed control plots. (After Towne et al. 2005; photo-
graph courtesy of Jim Peaco/US National Park Service.)

The most effectively managed parks are usually those whose


Management plans are
managers have the benefit of research and monitoring programs
needed that articulate
and have funds available to implement their management plans.
conservation goals
At these sites, the effects of different grazing methods (sheep
and practical methods versus cattle, light versus heavy) on populations of wildflowers,
for achieving them. butterflies, and birds are closely followed using an experimen-
Management activities tal approach in which treated areas are compared with control
can include controlled areas. These experiments demonstrate that parks often must be
burns, enforcement of actively managed to prevent deterioration.
restrictions on human
use, and the maintenance Managing sites
Introduction to Conservation Biology 1E Primack/Sher
of keystone resources, ParkAssociates
Sinauer managers sometimes must actively manage sites to ensure
especially water. that all
Morales or particular successional stages are present so that spe-
Studio/SA
Primack_Sher1E_08.15 Date 02-11-16 02-29-16 3-1-16
cies characteristic of each stage have a place to persist and thrive.
Protected Areas 291

In some wildlife sanctuaries, grasslands and fields are maintained by live-


stock grazing, burning, mowing, tree cutting, or shallow plowing in order
to retain open habitat in the landscape. One common way to do this is to
set localized, controlled fires periodically in grasslands, shrublands, and
forests to reinitiate the successional process (Middleton 2013). Obviously,
such burning must be done in a legal and carefully controlled manner to
prevent damage to nearby property (Figure 8.16A). Also, prior to burning,
land managers must develop a program of public education to explain to
local residents the role of fire in maintaining the balance of nature. In other
situations, parts of protected areas must be carefully managed to minimize
human disturbance and fire (Figure 8.16B).

(A)

Figure 8.16 Conservation man-


(B) agement: intervention versus leave-
it-alone. (A) Heathland in protected
areas of Cape Cod, Massachusetts,
is burned on a regular basis in order
to maintain the open vegetation
habitat and to protect wildflowers
and other rare species. (B) Some-
times management involves keeping
human disturbance to an absolute
minimum. Muir Woods National
Monument is a forest of old-growth
coast redwoods, protected in the
midst of the heavily urbanized San
Francisco Bay area. (A, photograph
by Elise Smith, US Fish and Wildlife
Service; B, photograph courtesy of
US National Park Service.)
292 Chapter 8

keystone resources In many protected areas, it may be necessary


to preserve, maintain, and supplement keystone resources on which many
species depend. These resources include trees that supply fruit when little
or no other food is available, pools of water during a dry season, exposed
mineral licks, and so forth. Keystone resources and keystone species can
be enhanced in managed conservation areas to increase the populations
of species whose numbers have declined.
By planting areas with food plants and building an artificial pond, it may
be possible to maintain vertebrate species in a smaller conservation area and
at higher densities than would be predicted based on studies of species dis-
tribution in undisturbed habitat. Artificial ponds, for example, can provide
habitat for insects that people enjoy watching such as dragonflies and also
serve as centers of public education in urban areas (Kobori and Primack
2003). Likewise, nest boxes or drilled nesting holes in living and dead trees
can provide substitute shelter resources for birds and mammals where there
are few dead trees with nesting cavities (Lindenmayer et al. 2009).
In each case a balance must be struck between, at one extreme, establish-
ing nature reserves free from human influence and, at the other extreme,
creating seminatural gardens in which the plants and animals are dependent
on people. Further, in some cases human replacements or augmentations of
resources may be ineffective, such as placing next boxes on small trees that
lack other features hollow-nesting birds need (Le Roux et al. 2015).
Rivers, lakes, swamps, estuaries, ponds, lakes, and all the other types
of wetlands must receive a sufficient supply of clean water to maintain
ecosystem processes. In particular, maintaining healthy wetlands is neces-
sary for populations of waterbirds, fish, amphibians, aquatic plants, and
a host of other species (Jähnig et al. 2011). Yet protected areas may end up
directly competing for water resources with agricultural irrigation proj-
ects, demands for residential and industrial water supplies, flood control
schemes, and hydroelectric dams. Wetlands are often interconnected, so
a decision affecting water levels and quality in one place will have rami-
fications for other areas. In particular, the construction of dams on major
rivers often completely alters the environmental conditions, eliminating
or reducing the abundance of many of the native fish and other aquatic
species. Consequently, one strategy for maintaining wetlands is to include
entire watersheds within given protected areas.

Monitoring sites
An important aspect of managing protected areas involves monitoring
components that are crucial for biodiversity, such as the quality and quan-
tity of water in ponds and streams; the number of individuals of rare and
endangered species; and the density of herbs, shrubs, and trees (Linden-
mayer et al. 2011; Pocock et al. 2015). Methods for monitoring these compo-
nents include recording standard observations, carrying out surveys, and
taking photographs from fixed points. Monitoring an area’s biodiversity
is sometimes combined with monitoring social and economic aspects of
surrounding communities because of the linkages between people and
Protected Areas 293

conservation. In particular, it is often important to monitor the amount


and value of plant and animal materials that people obtain from nearby
ecosystems.
Managers must continually assess the information they gain from
monitoring and adjust park management practices in an adaptive manner
to achieve their conservation objectives, a process sometimes referred to
as adaptive management (Figure 8.17).
Unfortunately, monitoring programs are often discontinued
after a few years from a lack of funding or interest. The lack of
long-term monitoring and baseline inventories can make it dif- Parks must be monitored
ficult or impossible to design and evaluate management plans to determine whether
and detect the effects of relatively subtle, chronic problems, such their goals are being met,
as acid rain and climate change, which can dramatically alter and management plans
ecosystems over time. To remedy this situation, many scientific
may need to be adjusted
research organizations and government agencies have begun to
based on new information
implement programs to monitor and study ecological change
from monitoring.
over the course of decades and centuries. Programs include the
system of 26 Long-Term Ecological Research (LTER) sites estab-
lished by the US National Science Foundation, and the World

Implement and act Monitor and


on management plan review

Develop Concerns based on


management plan biological and legal issues,
Protected area values, and expectations

Managers
Open, informed
debate Information
Scientists
gathering

General public,
government officials,
and others

Biology of Social, economic,


endangered species legal, and other
and habitats considerations

Figure 8.17 Model of an adaptive management process for protected


areas, emphasizing the decision-making stages. Input is solicited from many
sources, and then the plan is developed, implemented, and monitored. (After
Cork et al. 2000.)
294 Chapter 8

Network of Biosphere Reserves of the UN’s Educational, Scientific and


Cultural Organization (UNESCO).
The scale and methods of monitoring have to be appropriate for man-
agement needs. For large parks in remote areas, remote sensing using
satellites and airplanes may be an effective method for monitoring logging,
shifting cultivation, mining, and other activities, both authorized and un-
authorized (Morgan et al. 2010). In some cases, the inhabitants are trained
to carry out more intensive monitoring at a smaller scale. Often these local
people have extensive and useful knowledge of a protected area that they
are willing to share as part of the monitoring process (Anadón et al. 2009).

Management and people


A central part of any park’s management plan must be policies
The involvement of local on the use of the park resources by different groups of people
people is often the crucial (Grumbine and Xu 2011). In developing countries especially, re-
element missing from stricting access to protected areas can cost people—who may
conservation strategies.
have traditionally used the area and its resources—access to the
basic resources that they need to stay alive. Such displaced peo-
Local people need to be
ple may oppose conservation in these areas (Mascia and Claus
involved in conservation
2009). Many parks flourish or are destroyed depending on the
programs, as partici-
degree of support, neglect, or hostility they receive from the
pants, employees, and
people who live in or near them. In the best-case scenario, local
leaders. people become involved in park management and planning, are
trained and employed by the park authority, and benefit from
the protection of biodiversity and regulation of activity within
the park (Ferro et al. 2011).
However, in the worst-case scenario, if there is a history of bad relations
and mistrust between local people and the government or if the purpose
of the park is not explained adequately, the local people may reject the
park’s concept and ignore its regulations. In this case, the local people
may come into conflict with the park personnel, to the detriment of the
conservation goals. Park personnel and even armed soldiers may have to
patrol constantly to prevent illegal activity. Escalating cycles of conflict can
lead to outright conflict in which park personnel and the local inhabitants
may be threatened, injured, or even killed.
Unfortunately, it is sometimes necessary to exclude local people from
protected areas, especially in cases when resources are being overharvest-
ed, either legally or illegally, to the point that the health of the ecosystem
and the existence of endangered species are threatened (Packer et al. 2013).
Such degradation and loss of biodiversity can result from overgrazing by
cattle, excessive collection of fuelwood, or hunting with guns. In such cases
the only solution may be a strategy of “fences and fines.”

Zoning as a solution to conflicting demands


A possible way to deal with conflicting demands on a protected area is
zoning, which considers the overall management objectives for a park and
sets aside designated areas that permit or give priority to certain activi-
Protected Areas 295

ties. For example, some areas of a forest may be designated for


timber production, hunting, wildlife protection, nature trails, or Zoning allows the
watershed maintenance. Other zones may be established for the separation of mutually
recovery of endangered species, restoration of degraded commu- incompatible activities.
nities, and scientific research. The challenge in zoning is to find MPAs are often zoned
a compromise that people are willing to accept which provides with no-fishing areas
for the long-term, sustainable use of natural resources. Manag- where fish and other
ers often need to spend considerable effort informing the public marine organisms can
about what activities are acceptable in particular areas of a park recover from harvesting.
and then enforcing park regulations (Andersson et al. 2007).
For example, an MPA might allow
fishing in certain areas and strictly pro-
hibit it in others; certain areas might be
designated for surfing, water-skiing,
and recreational diving, but these
sports may be prohibited elsewhere. (A) Philippines
The creation of zoned MPAs in many
locations has proven to be an effective
Village
way to rebuild and maintain popula- Lighthouse
tions of fish and other marine organ-
isms (Figure 8.18). In comparison with
nearby unprotected sites, MPAs that re- Village
strict fishing and other activities often 50 × 20 m Apo
have greater total weight of commer- census Village Island
area

500 m
Figure 8.18
Fished nonreserve

On Apo Island in the


Philippines, large reef fish had been over-
harvested to the point where they were
rarely seen. (A) In response to overhar-
vesting, a reserve was set up (blue area) Unfished reserve
on the eastern side of the island, while
fishing continued as before at a specified
nonreserve area on the western side. A (B)
censusing study measured the number 25
of large reef fish at each site (six under- Census inside reserve
Mean density of large reef fish

water census areas are shown as black 20 Census outside reserve


(individuals/1000 m2)

rectangles for each site). (B) The resulting


Fish rapidly
data show that after the marine reserve increasing inside
was established, the number of fish ob- 15
reserve
served in the unfished reserve increased
substantially. The number of fish in the 10
unprotected area did not increase initially Fish slowly increasing
outside reserve
because the fish were still being inten-
sively harvested; after about 8 years, 5
however, an increase became detectable,
originating from the spillover of fish from
the reserve area. (After Abesamis and 0 3 6 9 12 15 18 21
Russ 2005.) Number of years of protection
296 Chapter 8

cially important fish, greater numbers of individual fish, and greater coral
reef cover. Evidence shows that fish from MPAs that restrict or prohibit
fishing spill over into adjacent unprotected areas, where they can help
rebuild populations and also may be caught by fishermen. The enforce-
ment of zoning is often a major challenge in MPAs because fishermen
tend to fish on the edges of fishing-exclusion zones as those are the areas
where the fishing is best, leading to overfishing at the margins of MPAs.
A combination of local involvement, publicity, education, clear posting of
warning signs, and visible enforcement significantly increases the success
of any zoning plan, especially in the marine environment (Fox et al. 2012).

Biosphere reserves
UNESCO has pioneered another zoning approach, termed biosphere re-
serves, under its World Network of Biosphere Reserves Program, which
integrates traditional land-use patterns (such as farming, grazing, and
managing forests), research, protection of the natural environment, and
sometimes tourism at a single location. These locations often have well-
established human settlements and scenic landscapes. A desirable feature
of the biosphere reserve program is a system in which there are zones
delineating varying levels of use (Figure 8.19A). At the center is a core
area in which ecosystems are strictly protected, with all human activity
either prohibited or tightly regulated. This core is surrounded by a buffer
zone in which traditional human activities, such as the collection of edible
plants and small fuelwood, are monitored and nondestructive research
is conducted. Surrounding the buffer zone is a transition zone in which
some forms of sustainable development, such as small-scale farming, are
allowed. In addition, some extraction of natural resources, such as selective
logging, and experimental research are also permitted. This general strat-
egy of surrounding core conservation areas with buffer and transitional
zones can encourage local people to support the goals of the protected area.
However, although these zones are easy to draw on paper, in practice it
has been difficult to inform and gain agreement from residents who live
in or near biosphere reserves about where the zones are and what uses
are allowed in them.
The value of the strategy of surrounding core conservation areas with
buffer and transition zones is still being debated. The approach has ben-
efits: local people may be more willing to support park activities if they
are allowed zoned access to the park, and certain desirable features of
the landscape created by human use may be maintained (such as farms,
gardens, and early stages of succession). Also, buffer zones may prevent
parks from becoming isolated islands of nature and may create corridors
that facilitate animal dispersal between highly protected core conservation
areas. Yet zoning for multiple uses and resource extraction may only work
if the core area is large enough to protect viable populations of all key spe-
cies and if people are willing to respect the zones and their designated uses.
Protected Areas 297

Respect for zones varies greatly in different parts of the world and among
different social situations. In places where park management, political will,
and land tenure are weak, buffer zones often are seen as a commons or
as unowned and unmanaged lands that are up for grabs, which greatly
reduces their effectiveness.
One instructive example of a biosphere reserve is the Kuna Yala Indig-
enous Reserve on the northeast coast of Panama. In this protected area
comprising 60,000 ha of tropical forest and coral islands live 50,000 Kuna
people, in 60 villages, who practice traditional medicine, agriculture, fish-
ing, and forestry while documentation and research are undertaken by
scientists from outside institutions (Figure 8.19B). At present, Kuna con-

(A)

Buffer Transition M Monitoring


Core area zone zone
M T Tourism and recreation
M M R
T Human settlements
R Research station,
T R education, training

(B)

Figure 8.19 (A) The general pattern of a biosphere reserve: a core protected
area is surrounded by a buffer zone, where human activities are monitored and
managed and where research is carried out; this, in turn, is surrounded by a
transition zone, where sustainable development and experimental research take
place. (B) The Kuna people still practice traditional methods of catching fish in
Kuna Yala Indigenous Reserve. (Photograph © Andoni Canela/AGE Fotostock.)
298 Chapter 8

servation beliefs and practices are gradually changing because of outside


influences, and younger Kuna often question the need to rigidly protect
the reserve. Further, rising sea levels and declining marine resources are
forcing village leaders to consider other options for their future (Posey
and Balick 2006).

Challenges to Park Management


Managers of protected areas face many challenges. Perhaps the most basic
and unavoidable of these challenges are increased pressures from growing
human populations and demand for the use of natural resources. Although
all threats to species discussed in Chapter 4 are relevant to protected areas,
the following are some areas of particular concern.

Poaching
Human populations will continue to increase dramatically in the coming
decades, while resources such as fuelwood, medicinal plants, and wild
meat will become harder to find. Similarly, people who are poor and hun-
gry will enter the nearby protected areas to take what they need to live,
regardless of whether they have permission (see “Overexploitation” in
Chapter 4). Within protected areas, if park rangers are underpaid, even
they may be motivated to begin illegally harvesting and selling the very
resources that they are charged with protecting. Addressing poverty and
enforcing regulations are the most important factors to address poaching
(see Chapter 12).

Trophy hunting
There is considerable debate about the role of hunting in conservation. On
one side, many countries finance conservation primarily from the sale of
hunting licenses. For developing countries, the focus is on large game such
as the hunting of rhinos or buffalo by wealthy foreigners (see Chapter 3).
Trophy hunting, when strictly regulated, has minimal effects on the overall
number of individuals within a park and can even double as a manage-
ment tool. One example of this is selling expensive licenses to hunt specific,
individual old male rhinos that are killing young rhinos and preventing
younger males from mating. However, critics argue that allowing sport
hunting of any type not only sends the wrong message, that killing these
animals is acceptable, but also supports a market that, in turn, promotes
illegal trophy hunting. This is illustrated by the killing of Cecil the Lion
(see Chapter 3 opener); because South Africa permits lion hunting, the
American hunter may not have known that the lion he was shooting had
been illegally taken from one of the protected areas. The financial stakes
are high enough to motivate local hunting guides to steal from the parks.
Making all trophy hunting illegal and removing these animal products
from circulation, as was done when the Kenyan government burned tons
of ivory, has helped in some cases (see Figure 6.24).
Protected Areas 299

Human–animal conflict
Problems are inevitable as more people live and farm closer to high con-
centrations of wildlife that, when food is scarce, have nowhere to go but
out of the park and into nearby agricultural fields and villages. Elephants,
primates, and flocks of birds can all be significant crop raiders, while car-
nivores such as tigers pose a different set of challenges to nearby residents.
Some nonprofit organizations and governments address these problems
by creating opportunities for local people to also benefit from the animals,
compensating them for their losses, and helping to build fences or other
deterrents (Figure 8.20).

Degradation
Multiple-use areas can suffer from the negative effects of mining, cattle
grazing, and oil exploration due to the lack of management or poor enforce-
ment of policies. Even strictly protected areas are at risk of degradation
from recreation, including wildfires, littering, fragmentation and erosion
from off-road driving, and the habituation of wild animals, to name just a
few (see Chapter 3). Furthermore, strictly protected areas may suffer from
the same degradation as multiple-use areas, including instances when lack
of oversight leads to illegal use. Generally speaking, degradation from
human use is minimized in multiple-use areas where there is oversight
and policies that regulate it.

Climate change
The extent to which existing protected areas will allow species and eco-
systems to persist in the face of climate change is an important question

Figure 8.20 Elephants from


protected areas trample locals
farms and lower support for their
conservation. In response, beehive
boxes established along the perim-
eters of farms create anti-elephant
“fences.” (© Steve Taylor ARPS/
Alamy Stock Photo.)
300 Chapter 8

and is being investigated (Regos et al. 2016). Species may not be able to
persist in a protected area if the climate or the associated vegetation change
significantly. For example, the Doñana wetlands in southern Spain are a
World Heritage site that is already rated as under “very high threat” by
UNESCO (Scheffer et al. 2015). The site contains some of the most impor-
tant overwintering habitat in Europe for migrating waterbirds, yet these
wetlands may dry out in coming decades due to the development of a
warmer, drier climate. Similarly, in one study of the Yunnan Province in
China, it was determined that as much as 65% of the region is expected to
shift to a different climate zone by 2050, shifting 83% of the total protected
area to a different bioclimactic stratum, to which the local species may not
be able to adapt in time (Zomer et al. 2015). It may be necessary to estab-
lish new protected areas in places where a species or ecosystem may be
able to disperse and survive in coming decades. In this rapidly changing
environment, it is important to preserve elevational and environmental
gradients, corridors, and climatic refugia so that species and ecosystems
can gradually spread in response to a changing climate.
Given that species ranges are already shifting and are predicted to
continue to do so in response to the rapidly changing climate, some con-
servationists have begun prioritizing areas for protection based on the
geophysical environment rather than the species that occur there; that is,
focusing on the stage (geology, soils, topography, etc.) rather than the actors
(species) (Groves et al. 2012). Biodiversity often corresponds to an area’s
geophysical diversity, although the strength of this association varies from
region to region (Anderson and Ferree 2010). The use of this strategy will
likely continue because data on geophysical characteristics are available for
most locations and the approach is relatively easy to integrate into existing
systems for identifying high-priority conservation areas.

Funding and personnel


For park management to be effective, there must be adequate funding for
a sufficient number of well-equipped, properly trained, reasonably paid,
and highly motivated park personnel who are willing to carry out park
policy (Becker et al. 2013; Tranquilli et al. 2012). Buildings, communications
equipment, and other appropriate elements of infrastructure are necessary
to manage a park. In many areas of the world, particularly in developing,
but also in developed, countries, protected areas are understaffed, and
the park staff lack the equipment to patrol remote areas. Without enough
radios and vehicles, they may be restricted to the vicinity of headquarters
and unaware of what is happening in their own park.
The importance of sufficient personnel and equipment should not be
underestimated: in areas of Panama, for instance, a greater frequency of
antipoaching patrols by park guards results in a greater abundance of
large mammals and the seed dispersal services they provide (Brodie et
al. 2009; Wright et al. 2000). International conservation organizations and
governmental agencies regularly assist in providing funds for managing
Protected Areas 301

protected areas in developing countries, but often the funding remains


insufficient. Increasing funding for the management of protected areas
needs to be a priority for government agencies and conservation organiza-
tions (Watson et al. 2014).
Finally, at the end of the day, conservation biologists need to account
for whether their management of protected areas achieved the stated goals
and whether the funds were spent effectively. As human impact increases
in most areas of the world, the importance of protected areas for the protec-
tion of biodiversity is expected only to increase (Regos et al. 2016).

Summary
„„Protecting habitat is the most effec- place and may be particularly important
tive method of preserving biodiversity. in maintaining known migration routes.
About 15.4% of the Earth’s land surface „„Protected areas often must be actively
is included in about 209,000 protected managed in order to maintain their biodi-
areas, but because of the needs of human versity. Monitoring provides information
societies for natural resources, the per- that is needed to evaluate whether man-
centage may not increase much further. agement activities are achieving their in-
„„Conservation biologists are developing tended objectives or need to be adjusted.
guidelines for designing protected areas: „„Management might involve zoning to
the areas should be large whenever pos- establish areas where certain activities
sible, they should not be fragmented, are allowed or prohibited. Managing
and managers should create networks interactions with local people is criti-
of conservation areas for maximum cal to the success of protected areas and
protection. should be part of a management plan.
„„Habitat corridors connecting protected „„Adequate staffing and funding are nec-
areas may allow species dispersal to take essary for park management.

For Discussion
1. Obtain a map of a town, state, or na- 3. Think about a national park or nature
tion that shows protected areas (such as reserve you have visited. In what ways
nature reserves and parks) and multiple- was it well run or poorly run? What were
use managed areas. Who is responsible the goals of this protected area, and how
for each parcel of land, and what is the could they be achieved through better
goal in managing it? Consider the same management?
issues for aquatic habitats (ponds, lakes, 4. Can you think of special challenges in the
rivers, coastal zones, etc.). management of aquatic preserves such as
2. If you could protect additional areas on coastal estuaries, islands, or freshwater
the map, where would they be and why? lakes that would not be faced by manag-
Show their exact locations, sizes, and ers of terrestrial protected areas?
shapes and justify your choices.
302 Chapter 8

Suggested Readings
Burgess, S. C., K. J. Nickols, C. D. Griesmer, L. A. K. Barnett, A. G. Dedrick, E.
V. Satterthwaite, and 4 others. 2014. Beyond connectivity: How empirical
methods can quantify population persistence to improve marine protect-
ed-area design. Ecological Applications 24: 257–270. Designers of a network
of MPAs should consider the dispersal ability of species.
Burrell, J. 2013. Path of the Pronghorn—Leading to New Passages: Part 3.
Newswatch, National Geographic. Assessment of fencing, overpasses,
and underpasses constructed to protect pronghorn migration to and
from Grand Teton National Park. (newswatch.nationalgeographic.
com/2013/12/06/path-of-the-pronghorn-leading-to-new-passages-part-3).
Colwell, R., S. Avery, J. Berger, G. E. Davis, H. Hamilton, T. Lovejoy, and 6
others. 2012. Revisiting Leopold: Resource Stewardship in the National Parks.
National Park System Advisory Board, Washington, DC. Protected area
management needs to consider that the environment is changing.
Danielson, F., P. M. Jensen, N. D. Burgess, R. Altamirano, P. A. Alviola, H. An-
drianandrasana, and 21 others. 2014. A multicountry assessment of tropical
resource monitoring by local communities. BioScience 64: 236–251. Moni-
toring by local people and by scientists produces similar results.
Edgar, G. J.R. D. Stuart-Smith, T. J. Willis, S. Kininmonth, S. C. Baker, S. Banks,
and 19 others. 2014. Global conservation outcomes depend on marine pro-
tected areas with five key features. Nature 506: 216–220. The characteristics
of successful MPAs are having no-take zones, having good enforcement,
being large, being old and established, and being isolated from areas with
fishing.
Hallwass, G., P. F. Lopes, A. A. Juras, and R. A. M. Silvano. 2013. Fishers’
knowledge identifies environmental changes in fish abundance trends in
impounded tropical rivers. Ecological Applications 23: 392–407. Local people
can sometimes accurately describe the changes in species composition and
abundance that have occurred after a dam has been built.
Hobbs, R. J., D. N. Cole, L. Yung, E. S. Zavaleta, G. H. Aplet, F. S. Chapin III,
and 10 others. 2010. Guiding concepts for park and wilderness steward-
ship in an era of global environmental change. Frontiers in Ecology and the
Environment 8: 483–490. Excellent statement of the need for guiding prin-
ciples in park management.
Joppa, L. N., P. Visconti, C. N. Jenkins, and S. L. Pimm. 2013. Achieving the
Convention on Biological Diversity’s goals for plant conservation. Science
341: 1100–1103. It is possible but will be difficult in practice to achieve the
goals of formally protecting 17% of the terrestrial world and 60% of plant
species.
Maron, M., J. R. Rhodes, and P. Gibbons. 2013. Calculating the benefit of con-
servation actions. Conservation Letters 6: 359–367. Most of the time, con-
servation benefits are calculated wrong, and this article offers improved
methods.
Mascia, M. B., S. Pallier, R. Krithivasan, V. Roshchanka, D. Burns, M. J. Mlotha,
and 2 others. 2014. Protected area downgrading, downsizing, and de-
gazettement (PADDD) in Africa, Asia, and Latin America and the Carib-
bean, 1900–2010. Biological Conservation 169: 355–361. PADD represents a
widespread and generally underappreciated threat to biodiversity.
Protected Areas 303

Packer, C., A. Loveridge, S. Canney, T. Caro, S. T. Garnett, M. Pfeifer, and 52


others. 2013. Conserving large carnivores: Dollars and fence. Conservation
Letters 16: 635–641. In a study of lions, sometimes fencing off a protected
area is the best and least expensive strategy to maintain populations.
Regos, A., M. D’Amen, N. Titeux, S. Herrando, A. Guisan, and L. Brotons.
2016. Predicting the future effectiveness of protected areas for bird conser-
vation in Mediterranean ecosystems under climate change and novel fire
regime scenarios. Diversity and Distributions, 22(1): 83–96. Models can be
used to manage future risk.
Skelly, D. K., S. R. Bolden, and L. K. Freidenburg. 2014. Experimental canopy
removal enhances diversity of vernal pool amphibians. Ecological Applica-
tions 24: 340–345. Limited tree removal is an effective strategy for encour-
aging certain amphibian populations.
Vidal, O., J. López-García, and E. Rendón-Salinas. 2014. Trends in deforesta-
tion and forest degradation after a decade of monitoring in the Monarch
Butterfly Biosphere Reserve in Mexico. Conservation Biology 28: 177–186.
Management actions have been effective at reducing deforestation and
protecting butterfly habitat.
Wuerthner, G., E. Crist, and T. Butler (eds.). 2015. Protecting the Wild: Parks and
Wilderness. The Foundation For Conservation. Island Press. Researchers
from around the world make a strong case for the importance of protected
areas for conservation.
Conservation Outside
9 Protected Areas
The Value of Unprotected Working with Local People 321
Habitat 307 Case Studies: Namibia and
Conservation in Urban and Other Kenya 330
Human-Dominated Areas 313
Ecosystem Management 318

The black rhino (Diceros bicornis) has benefited from conservation


efforts outside the boundaries of national parks. At Imire Game Farm in
Zimbabwe, rhinos are protected from poaching by 24-hour armed guards.
E
stablishing protected areas with intact eco-
systems is essential for species conservation.
It is, however, shortsighted to rely solely on
protected areas to preserve biodiversity. That kind of
reliance can create a paradoxical situation in which
species and ecosystems inside the protected areas
are preserved while the same species and ecosystems
outside are allowed to be damaged, which in turn re-
sults in the decline of biodiversity within the protect-
ed areas (Kinnaird and O’Brien 2013; Newmark 2008).
This decline is due to at least three reasons:

• Most protected areas are too small to protect viable,


long-term populations of many species, especially
large animals and migratory species.
• Many species are attracted to resources available out-
side protected areas. For example, it is not uncommon
for primate species to raid crops in villages that are
adjacent to protected areas in which they live.
• Many species migrate between protected areas sea-
sonally to avoid freezing temperatures or other climate
extremes or to access mates and other resources.
In general, the smaller a protected area, the more
dependent it is on neighboring unprotected lands for
the long-term maintenance of biodiversity. A crucial
component of conservation strategies must be the
306 Chapter 9

protection of biodiversity inside and outside—both


immediately adjacent to and away from—protect-
ed areas (Troupin and Carmel 2014). According to
even the most optimistic predictions, in the future
more than 80% of the world’s land will remain out-
side protected areas (PAs). Given this reality, the
UN Convention on Biological Diversity (see Chap-
ter 11) established global conservation targets that
include both PAs and what they call “other effec-
tive area-based conservation measures” (OECM’s;
Aichi Target 11). Conservation biologists must help
define “effective conservation” and determine to
what degree OECM’s contribute to global targets
Figure 9.1 Planting a variety of species in (Watson et al. 2016).
urban parks and gardens is one way in which Some conservation biologists argue that, given
landscapes dominated by humans can support how much land is now occupied by people, we must
biodiversity. Many insects, birds, and other find ways to promote biodiversity in human-domi-
taxa are dependent on these habitat stepping- nated landscapes (Figure 9.1). This is an idea ecolo-
stones to travel between protected areas. gist Michael Rosenzweig (2003) named reconcilia-
(© johandersonphoto/Getty Images.) tion ecology. In many cases, this involves creating
habitat in urban settings, such as enhancing parks
or planting green roofs (see chapter opening photo
for Chapter 11), and it generally involves increasing the environmental com-
plexity (Loke et al. 2015). An innovative example of reconciliation ecology
is an underwater restaurant in Eilat, Israel, which has frames built outside
the windows to support a coral nursery. The creation of new reef habitat on
the frames helps mitigate damage by snorkeling tourists on the main reef
nearby, while creating a beautiful view for diners at the restaurant. Other
examples of how nature is promoted in urban areas, and also some of the
problems of this approach, are discussed in more detail later in the chapter.
Jeff McNeely (1989), an International Union for Conservation of Nature
(IUCN) expert in protected areas, suggested that the park boundary
is too often also a psychological boundary, suggesting that since nature
is taken care of by the national park, we can abuse the surrounding
lands, isolating the national park as an ‘island’ of habitat which is
subject to the usual increased threats that go with insularity.

In the worst case, a devastated landscape polluting the air and


Many endangered species water will strangle the protected area it surrounds and block
and unique ecosystems the movement of dispersing animals and plants. In the best
are found partly or case, however, unprotected areas surrounding protected areas
entirely on unprotected will provide additional space for ecosystem processes and new
lands. Consequently, populations. In many ways, conservation outside protected
the conservation of bio- areas should strive to blur the distinctions between protected
diversity in these places and unprotected ecosystems as much as possible by maintain-
must be considered. ing unprotected areas in a state of reasonable ecological health
(Radeloff et al. 2010).
Conservation Outside Protected Areas 307

In this chapter, we explore strategies to include biodiversity protection


as a management objective for unprotected areas as a way to complement
conservation in protected areas.

The Value of Unprotected Habitat


The human use of ecosystems varies greatly in unprotected lands, but sig-
nificant portions are not used intensively by humans and still harbor some
of their original biota (Figure 9.2). In almost every country, numerous
Free

Wilderness

Multi-use
forest
Within

National park

Degraded
Low

wildlands
Historic range of variability

Rangeland
Dam density
Examples:

Traditional
Processes

agriculture

Exurban
Plantations
High

(intensive agriculture)
Far outside

Industrial Urban open space


and urban Suburban
Controlled

City park

Artificial Patterns Natural


Examples:
High Housing and road density Low
Low Proportion of natural vegetation cover High

Figure 9.2 Landscapes vary in the extent to which humans have altered the
patterns of species composition and natural vegetation through activities such
as agriculture, road construction, and housing; ecosystem processes (water flow,
nutrient cycling, etc.) also vary because of fire control activities, dam construc-
tion, and other activities that alter plant cover. Wilderness areas retain most of
their original patterns and processes, urban areas retain the least, and other
landscapes retain various intermediate amounts. (After Theobald 2004.)
308 Chapter 9

rare species and ecosystems exist primarily or exclusively on unprotected


public lands or on lands that are privately owned (Deguise and Kerr 2006).
In the United States, 60% of species that are globally rare or listed under
the US Endangered Species Act live on private forested lands (Robles et
al. 2008). Even when endangered species are found on a country’s public
land, it is often not land managed for biodiversity but rather land managed
primarily for timber harvesting, grazing, mining, or other economic uses.
For example, 75% of the remaining orangutans in Indonesia live outside
protected forests, often in logged forests and tree plantations (Meijaard et
al. 2010). Clearly, strategies for reconciling human needs and conservation
interests in unprotected areas are critical to the success of conservation
plans (Cox and Underwood 2011; Koh et al. 2009).
The situation of the Florida panther (Felis [Puma] concolor coryi) pro-
vides an excellent example of the importance of unprotected habitat. This
endangered subspecies of mountain lion lives in South Florida and has a
population of only 100–120 individuals (Thatcher et al. 2009). About 31%
of the land in the present range of the panther is privately owned, and
animals tracked with radio collars have all spent at least some of their time
on private lands (Figure 9.3). Private lands typically have better soils,
which support more prey species than the public lands. Thus, panthers
that spend most of their time on private lands have a better diet and are
in better condition. Although panther habitat was protected on 870,362 ha
of public lands as of 2013, much more habitat must be protected in order
to ensure the panther’s continued survival. The acquisition of additional
habitat has been proposed (Kautz et al. 2006), but until that can be achieved,
even slowing down the pace of land development may prove impractical.
Two viable possibilities are educating private landowners on the value of
conservation and paying willing landowners to practice management op-
tions that allow for the panthers’ continued existence—specifically, mini-
mizing habitat fragmentation and maintaining their preferred habitats of
hardwood forest and cypress swamp. In addition, special road underpasses
have been built in the hopes of reducing panther deaths from collisions
with motor vehicles. Other species of big cats also frequently live outside
protected areas; for example, leopards in India often adapt well to rural
landscapes of farms and villages, hunting wild prey at night, when they
are rarely or never seen by villagers (Athreya et al. 2013).
Next, we will discuss the importance of several types of land that are
not contained within traditional protected areas but nonetheless are im-
portant for biodiversity.

Military land
Native species can often continue to live in unprotected areas, especially
when those areas are set aside or managed for some other purpose that is
not harmful to the ecosystem, such as security zones surrounding govern-
ment installations and military reservations. For example, the US Depart-
ment of Defense manages more than 11 million ha, much of it undevel-
Conservation Outside Protected Areas 309

(A) Figure 9.3 (A) The Florida


panther is found on both public and
private lands in South Florida. (B)
The green dots represent 55,000
radio telemetry records of 79 collared
panthers. Public lands are outlined in
red (A, photograph courtesy of Larry
Richardson, USFWS; B, from Kautz et
al. 2006, with updates from R. Kautz.)

(B)

Florida
Tampa

Lake
Okeechobee

Caloosahatchee Palm Beach


River

50 km Fort Myers
Naples

Radio-telemetry record
Miami
Public lands
(as of September 2013)
Wetlands Everglades
National
Main study area Park

oped, containing about 420 threatened and endangered species of plants


and animals. The White Sands Missile Range in New Mexico alone is
almost 1 million ha in area, about the same size as Yellowstone National
Park. While certain sections of military reservations may be damaged by
military activities, much of the habitat remains as an undeveloped buf-
310 Chapter 9

fer zone with restricted access. After these areas are no longer needed
for military purposes, they also make excellent candidates for protected
areas; nearly 31,000 hectares of former military bases are being converted
to nature reserves by the German Federal Agency for Nature Conserva-
tion (Huffington Post 2015).
The impact of military training itself, including accidental fires, tank
exercises, and artillery practice, provides the open habitat required by
certain species, such as the Karner blue butterfly and its host plants near
Fort McCoy in Wisconsin. As a result, many military bases have become de
facto refuges for about 420 federally listed species of plants and animals,
many of which have their largest populations on military bases (Stein et al.
2008). Rare and endangered desert tortoises (Gopherus agassizii), manatees
(Trichechus manatus), red-cockaded woodpeckers (Picoides borealis), bald
eagles, Atlantic white cedars (Chamaecyparis thyoides), and the least Bell’s
vireo (Vireo bellii pusillus) all have found a safe haven on military lands.
Personnel at the Barksdale Air Force Base in Shreveport, Louisiana, have re-
flooded wetlands along the Red River, restoring wetlands for wading birds.
The US Department of Defense’s emphasis on conservation has increased
dramatically; spending on threatened and endangered species jumped
45%, from about $50 million in 2003 to $73 million in 2012 (Watson 2013).
On the other hand, many military bases contain toxic waste dumps and
high levels of chemical pollutants. In addition, severe disturbance in the
form of bomb explosions, artillery practice, and the use of heavy vehicles
can have significant negative effects on the resident wildlife.

Unprotected forests
Forests that are either selectively logged on a long cutting cycle or are cut
down for farming using traditional shifting cultivation methods may still
contain a considerable percentage of their original biota and maintain most
of their ecosystem services (Adum et al. 2013; MacKay et al. 2014) (Figure
9.4). This is particularly true if fires and erosion have not irreversibly dam-
aged the soil and if native species can migrate from nearby undisturbed
lands, such as steep hillsides, swamps, and river forests, and colonize the
sites. For example, in Malaysia, most forest bird species are still found
in rain forests 30 years after selective logging was carried out, and un-
disturbed forest is available nearby to act as a source of colonists (Peh et
al. 2005). Likewise, in African tropical forests, gorillas, chimpanzees, and
elephants can tolerate selective logging and other land uses that involve
low levels of disturbance, though only when hunting levels are controlled
by active antipoaching patrols (Stokes et al. 2010).

Unprotected grasslands
The mown edges of roadsides often provide an open grassland community
that is a critical resource for many species, such as butterflies (Saarinen et
al. 2005). A similar habitat is provided by the surprisingly large amount
of mown fields occupied by power lines. In the United States, corridors for
Conservation Outside Protected Areas 311

Native forest species


60
Open-habitat species
50 Not evaluated
Number of species

40

30

20

10

0
Baseline Logged Secondary Unused Tree Perennial Annual Clearcut
forest forest forest field plantation crops crops
Low impact Intensity of human impacts High impact

Figure 9.4 For a range of land uses in West Africa, when the intensity of hu-
man impacts increases, the average number of vertebrate native forest species
declines and the number of open-habitat species increases. Some native forest
species are still present even along with intensive land use, such as tree planta-
tions, but the overall number of species and the proportion of native species are
much lower. (After Norris et al. 2010.)

power line rights-of-way occupy over 2 million ha. Power line corridors
managed with infrequent mowing and without herbicides maintain high
densities of birds, insects, and other animals (King et al. 2009). If such man-
agement practices could be extended over a greater proportion of power
line rights-of-way, these areas could become additional habitat for insects
and a wide range of other species. Remnant prairies in the United States
also represent an important habitat for many grassland species, especially
where the prairies can be managed with grazing or burning.

Unprotected waters
Many heavily altered aquatic ecosystems can also have value for conserva-
tion. For example, in estuaries and seas managed for commercial fisheries,
many of the native species remain because commercial and noncommercial
species alike require an undamaged chemical and physical environment.
It has been determined that most marine plant diversity (such as sea grass
and mangroves) occurs outside existing MPAs (Daru and le Roux 2016).
Also, many marine animals such as salmon, whales, and sea turtles migrate
great distances, including across areas that are not protected.
Even though dams, reservoirs, canals, dredging operations, port facili-
ties, and coastal development harm native aquatic communities, some bird,
fish, and other aquatic species are capable of adapting to the altered condi-
tions, particularly if the water is not polluted. However, there is abundant

Introduction to Conservation Biology 1E Primack/Sher


Sinauer Associates
Morales Studio/SA
312 Chapter 9

research that suggests that even when there is no obvious pol-


Even ecosystems that lution, many species are likely to have higher abundance within
are managed primarily MPAs than outside of them (e.g., Pikesley et al. 2016).
for the production of
Land that is undesirable to humans
natural resources can
retain considerable Other areas that are not protected by law may retain species
biodiversity, and they are
because the human population density and degree of use are
typically very low. Border areas, such as the demilitarized zone
important to the success
between North Korea and South Korea, often have an abundance
of conservation efforts.
of wildlife because they remain undeveloped and generally un-
occupied by people. Governments frequently manage mountain
areas, which are often too steep and inaccessible for develop-
ment, as valuable watersheds that produce a steady supply of water and
prevent flash flooding and erosion. They also harbor important natural
communities. Likewise, desert and tundra species and ecosystems may be
at less risk than other unprotected communities because such regions are
marginal for human habitation and use (MEA 2005). However, it is also
true that in areas like the Arctic region, the current warming of the climate
will result in further development of the transportation infrastructure and
a greater interest in mining deposits of oil, gas, and minerals.

Private land
In many parts of the world, wealthy individuals have acquired large tracts
of land for their personal estates and for private hunting. These estates
are frequently used at very low intensity, often in a deliberate attempt
by the landowners to maintain large wildlife populations. In particular,
some estates in Europe preserve unique old-growth forests that have been
owned and protected for hundreds of years by royal families. Such pri-
vately owned lands, whether owned by individuals, families, corporations,
or tribal groups, often contain important aspects of biodiversity.
Management for biodiversity can vary a great deal between landowners,
of course; a study of private landowners enrolled in the Indiana Classified
Forest and Wildlands Program found that size, environmental motives,
and those landowners who had seen improvements occur on their land
were more likely to be good stewards (Farmer et al. 2016). Strategies that
encourage private landowners and government land managers to protect
rare species and ecosystems are obviously essential to the long-term conser-
vation of biodiversity. This chapter and Chapter 12 describe these strategies.
Even small yards and home gardens can be useful for supporting bio-
diversity, particularly of insects (Ribeiro et al. 2016). The National Wild-
life Federation has a backyard “wildlife certification program” in which
homeowners can receive a certificate and a sign once they ensure that
their property contains all the elements of wildlife habitat, including a
food source, water, and sheltering plant cover for protection and reproduc-
tion. Homeowners’ associations may require the use of native plants or
Conservation Outside Protected Areas 313

(A) (B)

Figure 9.5 (A) This housing development in Lake Worth, Florida has wet-
land conservation areas that are managed largely by the local homeowners’
association. The creation of areas such as this are one way that developers can
legally mitigate their negative impact on the habitat that was displaced, while
also increasing the value of the houses they build. (B) The importance of these
conservation areas is apparent by the diversity of plants and animals that live
there, often appearing in people’s yards, like this sandhill crane (Grus canaden-
sis). Many sandhill cranes migrate from protected areas in the north each winter,
and are dependent on such remnants of habitat, while others are year-round
residents. (Photographs by Anna Sher.)

a minimum number of trees in landscaping to support biodiversity, and


developments may even have homeowners’-association-supervised natural
areas that increase the value of their properties (Figure 9.5).

Conservation in Urban and Other


Human-Dominated Areas
Many native species can persist even in urban areas, in public parks,
streams, ponds, and other, less altered, habitats (Meffert and Dziock 2012).
Preserving these remnants of biodiversity within a human-dominated ma-
trix not only presents special challenges but also provides unique opportu-
nities to educate the public about biodiversity conservation. For example,
the discovery of a new species of salamander at a popular swimming site
in Austin, Texas, required a change in how the site is managed to allow the
people and the salamanders to coexist. Likewise, in Europe, storks (Ciconia
ciconia) often nest in chimneys and towers, and endangered raptors such as
the peregrine falcon (Falco peregrinus) and bald eagle (Haliaeetus leucocepha-
lus) make nests and raise their young in the skyscrapers of downtown New
York, where numerous small animals (including the ubiquitous pigeons
314 Chapter 9

Figure 9.6 ”Pale Male” (on the left)


is a famous red-tailed hawk that has lived
on a Fifth Avenue residence in New York
city since 1991. The unusually light-
colored hawk has had eight mates thus
far and several broods. He has inspired
a website (www.palemale.com), at least
three children’s books, and even a movie.
He is one of the first hawks known to
have built a nest on a building in this city,
and when the homeowners attempted to
displace the hawks by removing the sup-
port structures for the nest, local bird lov-
ers successfully protested to keep it. Pale
Male is seen here with his then-current
mate, Lola. (© D. Bruce Yolton.)

and rats common to urban centers) provide abundant food sources (Figure
9.6). Even ponds at golf courses in urban areas and gravel pits dug for con-
struction materials may be suitable habitats for certain newts, dragonflies,
and other wetland species provided the water is not polluted (Colding et
al. 2009). In one study of 27 artificial water bodies in Australia, researchers
found that greater than 70% of the regional diversity of fish species could
be found in these human constructions (Davis and Moore 2015). Whether
intentional or not, these are examples of reconciliation ecology because
they demonstrate ways in which humans and other species can coexist.
As exciting as such examples of urban adaptations might be, we cannot
assume that all species have the potential to live within human-dominated
landscapes. For example, the value of urban parks for biodiversity found
in some developed countries may not apply to rapidly growing megacities;
in South America, they were found to be dominated by European weeds
(Fischer et al. 2016). We have a lot to learn about just what habitat and
disturbance features are important for various species and how to inte-
grate those features into our urban and suburban landscapes. In general,
increasing the intensity of land use will decrease the number of native spe-
cies found in a location, and adaptable, generalist species (often nonnative
invasives) will tend to do best. The size and configuration of landscape
features will determine which species and ecosystem processes are main-
tained. For example, abandoned industrial sites in Germany of at least 5
ha in area are necessary to provide habitat for many bird species of special
conservation concern (Meffert and Dziock 2012). More work is needed to
evaluate how general conservation principles apply in specific locations.
Increasing the presence of wild animals in the urban landscape comes
with fairly serious consequences for both animals and humans. For ex-
ample, as woodland areas and mountain canyons become urbanized or
Conservation Outside Protected Areas 315

suburbanized, people tend to create yards and gardens that attract deer.
Deer bring with them a host of problems: they can carry ticks that transmit
illnesses to humans, such as Lyme disease and Rocky Mountain spot-
ted fever; they are a significant potential road hazard; and the bucks can
become aggressive toward humans during mating season. In some areas,
deer that live within developments also attract predators including cougars,
thus increasing the potential for human–wildlife conflicts for a scarce and
ecologically important top carnivore.
Understanding the ecology, the ecosystem processes, and the char-
acteristics of the human use of a location is critical for implementing
policies to promote conservation in unprotected urban areas. Deciding
on the proper tools, though, requires good information on ecology and
complex urban human–natural systems and knowledge of how best to
motivate people to behave in conservation-friendly ways. These areas of
research are growing and beginning to provide insights that are improv-
ing urban conservation.

Other human-dominated landscapes


Most of the world’s landscapes have been affected in some way by human
activity, but fortunately, considerable biodiversity can be maintained in
well-managed and low-intensity traditional agricultural systems, grazing
lands, hunting preserves, forest plantations, and recreational lands (Car-
rière et al. 2013; Wright et al. 2012). Birds, insects, and other animal and
plant species are often abundant in traditional agricultural landscapes,
with their mixture of small fields, hedges, and woodlands. Some species
are found almost exclusively in these traditional human-dominated habi-
tats. In comparison with more intensive, so-called modern, agricultural
practices, which emphasize high yields of crops for sale in the market,
mechanization, and external inputs, these traditional landscapes experi-
ence less exposure to herbicides, fertilizers, and pesticides and have more
heterogeneity of habitat. Similarly, farmlands worked using organic meth-
ods support more birds than farmlands worked using nonorganic meth-
ods, in part because organic farms have more insects for the birds to eat.
In many areas of the world, however, the best agricultural lands are being
more intensively used while less optimal lands are abandoned as people
leave for urban areas (Phelps et al. 2013).
Conservation biologists are increasingly discussing the value of the
strategy of land sharing, in which low-intensity human activities, such
as traditional or organic agriculture, can coexist with some elements of
biodiversity. The alternative is land sparing, in which intensive human
activities, such as modern agriculture, are practiced on some of the lands
while allowing the rest to remain in their natural state. The best strategy
for any given location will depend on the local circumstances, the price
of land and crops, and potential financial incentives (Baudron and Giller
2014; Tscharntke et al. 2012).
316 Chapter 9

Figure 9.7 Two types of coffee (A)


management systems. (A) Shade
coffee is grown under a diverse
canopy of trees, providing a forest
structure in which birds, insects,
and other animals can live. (B) Sun
coffee is grown as a monoculture,
without shade trees. In a monocul-
ture system, animal life is greatly
reduced. (A, © Aurora Photos/Ala-
my; B, © PisitBurana/istock.)

(B)

One notable example of preserving biodiversity in an agricultural


setting comes from tropical countries and their traditional shade cof-
fee plantations, in which coffee is grown under a wide variety of shade
trees, with often as many as 40 tree species per farm (Philpott et al. 2007)
(Figure 9.7). In northern Latin America alone, shade coffee plantations
cover 2.7 million ha. These plantations have structural complexity created
by multiple vegetation layers and a diversity of birds and insects compa-
rable to the adjacent natural forest, and they represent a rich repository of
biodiversity (Vandermeer et al. 2010). The presence of such coffee planta-
tions can also potentially slow the pace of deforestation (Hylander et al.
2013). Therefore, programs are being developed to encourage and subsidize
farmers to maintain their shade-grown coffee plantations and to market
the product at a premium price as “environmentally friendly” shade-grown
coffee. But let the buyer beware; there are currently no uniform standards
for shade coffee. Thus, some coffee marketed as “environmentally friendly,
Conservation Outside Protected Areas 317

shade coffee” may actually be grown as sun coffee with only a few small,
interspersed trees. Shade-grown chocolate and other tropical tree crops
are similarly unregulated (Waldron et al. 2012).
In developing countries, conservation biologists have recently started
innovative programs in which local people living in rural areas are paid
directly for protecting individuals and populations of flagship species,
including rhinos, tigers, gorillas, and other species of conservation inter-
est (Dinerstein et al. 2013). When the animals do well, the people are paid
directly or receive money for village improvements (see the case studies
that follow).
In many countries, large parcels of government-owned land are des-
ignated as multiple-use habitat; that is, they are managed to provide a
variety of goods and services. An emerging and important research area
involves the development of innovative ways to reconcile competing claims
on land use, such as logging, mining, species conservation, and tourism.
This will require careful analyses and consideration of the trade-offs of
pursuing alternative development options in regard to both environmental
and socioeconomic priorities (Koh et al. 2010). A different approach is to use
regulations, the legal system, and political pressure to prevent government-
approved activities on public lands if these activities threaten the survival
of endangered species.
In the United States, the Bureau of Land Management oversees more
than 110 million ha, including 83% of the state of Nevada and large amounts
of Utah, Wyoming, Oregon, and Idaho. National forests cover over 83 mil-
lion ha, including much of the Rocky Mountains, the Cascade Range, the
Sierra Nevada, the Appalachian Mountains, and the southern coast of
Alaska. In the past, these lands have been managed for logging, mining,
grazing, wildlife, and recreation. The challenge is that often, each one of
these activities is managed by itself but their cumulative effects threaten
biodiversity. Increasingly, multiple-use lands also are being valued and
managed for their ability to protect species, biological communities, and
ecosystem services (Kemp et al. 2013). The US Endangered Species Act of
1973 and other similar laws, such as the 1976 National Forest Management
Act, require landowners, including government agencies, to avoid activities
that threaten listed species. One such activity is overgrazing by cattle; when
cattle grazing is reduced or eliminated on overgrazed rangelands, some of
these ecosystems can recover in a few years or decades (Earnst et al. 2012).
Another approach to protecting biodiversity in human-dominated
landscapes has been to define standards of best practices so that the use
of resources does not harm biodiversity. The Forest Stewardship Council
has been one such organization by working to promote the certification
of timber produced from sustainably managed forests. For the Forest
Stewardship Council and similar organizations to grant certification, the
forests need to be managed and monitored in the interests of their long-
term environmental health, and the rights and well-being of local people
and workers need to be protected. The certification of forests is increasing
rapidly in many areas of the world, especially in response to buyers in
318 Chapter 9

Europe, who often request certified wood products. At the same time,
major industrial organizations representing such industries as logging,
mining, and agriculture are lobbying for their own alternative certification
programs, which generally have lower requirements for monitoring and
weaker standards for judging whether practices are sustainable.

Ecosystem Management
Resource managers around the world are increasingly being
Ecosystem management urged by their governments and conservation organizations
links private and public to think on larger geographic scales, particularly given climate
landowners, businesses, change–driven shifts in the distributions of species and makeup
and conservation organ- of ecosystems. Traditionally, these managers may have focused
izations in a planning on the production of goods and services that could be managed
on the local scale, such as volume of timber or number of park
framework that facilitates
visitors. But today, these managers are being asked to expand
acting together on a large
their emphasis to a broader perspective that includes the conser-
scale.
vation of biodiversity and the protection of ecosystem processes
(Altman et al. 2011). That is, they are shifting to ecosystem man-
agement, a system of large-scale management involving multiple
stakeholders, the primary goal of which is preserving ecosystem
components and processes for the long term while still satisfying the current
needs of society (Figure 9.8). Rather than having each government agency,
private conservation organization, business, or landowner act in isolation
and in its own interests, ecosystem management envisions them cooperating
to achieve common objectives (Redpath et al. 2013). For example, in a large
forested watershed along a coast, ecosystem management would link all
owners and users located from the tops of the hills to the seashore, including
foresters, farmers, business groups, townspeople, and the fishing industry.
Important themes in ecosystem management include the following:
• Using the best science available to develop a coordinated plan for the
area that is sustainable; includes biological, economic, and social com-
ponents; and is shared by all levels of government as well as business
interests, conservation organizations, and private citizens
• Ensuring viable populations of all species, representative examples of
all biological communities and successional stages, and healthy eco-
system functions
• Seeking and understanding connections between all levels and scales in
the ecosystem hierarchy—from the individual organism to the species,
community, ecosystem, and even regional and global scales
• Monitoring significant components of the ecosystem (numbers of in-
dividuals of significant species, vegetation cover, water quality, etc.),
gathering the needed data, and then using the results to adjust man-
agement in an adaptive manner—a process sometimes referred to as
adaptive management (see Figure 8.17)
Conservation Outside Protected Areas 319

Mountain
recreation area and
water catchment
Wildlife refuge

Residential and
industrial area Forest plantation

Grazing
Fishing village
Sewage treatment plant

Farming
Wildlife refuge
Recreation and coastal No-fishing zone
development due to water
pollution

Coastal fishing
Marine protected area

Figure 9.8 Ecosystem management involves bringing together all the


stakeholders that affect a large ecosystem and receive benefits from it. In this
case, a watershed needs to be managed for a wide variety of purposes, many of
which influence one another. (After Miller 1996.)

One successful example of ecosystem management is the work of the


Malpai Borderlands Group, a nonprofit cooperative enterprise formed by
ranchers and other local landowners who promote collaboration among
conservation organizations such as The Nature Conservancy, private land-
owners, scientists, and government agencies (www.malpaiborderlands-
group.org). The group is developing a network of cooperation across the
Malpai planning area, which comprises nearly 400,000 ha of unique, rugged
mountain and desert habitat along the Arizona and New Mexico border.
The Malpai Borderlands Group uses controlled burning as a range man-
agement tool, reintroduces native grasses, applies innovative approaches
to cattle grazing, incorporates scientific research into management plans,
and takes action to avoid habitat fragmentation by using conservation
easements (agreements not to develop land) to prevent residential develop-

Introduction to Conservation Biology 1E Primack/Sher


Sinauer Associates
Morales Studio/SA
320 Chapter 9

ment. Their goal is to create “a healthy, unfragmented landscape to support


a diverse, flourishing community of human, plant and animal life in the
Borderlands Region” (Allen 2006).
A logical extension of ecosystem management is bioregional manage-
ment, which integrates protection with human use and often focuses on a
single large ecosystem, such as the Caribbean Sea or the Great Barrier Reef
of Australia, or on a series of linked ecosystems, such as the protected areas
of Central America. A bioregional approach is particularly appropriate
where there is a single, continuous, large ecosystem that crosses interna-
tional boundaries or when activity in one country or region will directly
affect an ecosystem in another country. For the European Union and the
21 individual countries that participate in the Mediterranean Action Plan
(MAP), for example, bioregional cooperation is absolutely necessary be-
cause the enclosed Mediterranean Sea has large human populations along
the coasts, heavy oil tanker traffic, and weak tides that cannot quickly
remove pollution resulting from cities, agriculture, and industry (Figure
9.9). This combination of problems threatens the health of the entire Medi-

France
Austria
Switzerland Hungary
Slovenia Romania
Croatia
Italy
Bosnia
Serbia Black Sea
Montenegro Bulgaria
Corsica Rome Macedonia Istanbul
Madrid Albania
Sardinia Greece
Turkey
Spain
Athens
Tunis Sicily

Mediterranean Sea Crete Cyprus


Algeria
Morocco Tripoli
Tunisia Tel Aviv
Alexandria Israel
Protected areas
Jordan
0–450 hectares Cairo
N
451–16,000 hectares
Libya Saudi
16,001–1,474,000 hectares Egypt Arabia

Figure 9.9 The countries participating in MAP cooperate in monitoring and


controlling pollution and coordinating the management of their protected areas.
Major protected areas along the coast are shown as dots. Note that there are no
major protected areas on the coasts of France, Libya, or Egypt due to private land
ownership and the promotion of economic development. (After Miller 1996.)
Conservation Outside Protected Areas 321

terranean ecosystem, including the sea, its surrounding lands, and its asso-
ciated tourist and fishing industries. Cross-boundary management is also
necessary because pollution from one country can significantly damage
the natural resources of neighboring countries. At a conference celebrating
MAP’s fortieth year, a representative of the United Nations Environmental
Program was quoted as saying, “While 2015 will be remembered as a major
milestone in terms of international agreements, 2016 has been called the
year for implementation and delivery. We must seize that opportunity, and
ride that momentum.” (UNEPMAP 2016.)

Working with Local People


Even remote regions that are considered “wilderness” by gov-
ernments and the general public often have small, sparse human In many parts of the
populations. Societies that practice a traditional way of life in world, areas with high
rural areas with relatively little outside influence in terms of biodiversity are inhabited
modern technology are variously referred to as “tribal people,” by indigenous people
“indigenous people,” “native people,” or more generally, “tradi- with long-standing
tional people” (Timmer and Juma 2005; www.iwgia.org). These systems for resource
people regard themselves as the original inhabitants or long- protection and use. These
standing residents of a region and are often organized at the
people are important,
community or village level.
and may be essential, to
It is necessary to distinguish these established traditional
conservation efforts in
peoples from more recent settlers, who may not be as concerned
those areas.
with the health of surrounding biological communities or as
knowledgeable about the species living there and the land’s
ecological limits. In many countries, such as India and Mexico,
there is a striking correspondence between areas occupied by
traditional people and the areas of high conservation value and intact forest
(Toledo 2001). Such local people often have established systems of rights to
natural resources, which sometimes are recognized by their governments,
and they are potentially important partners in conservation efforts (Rai
and Bawa 2013).
Worldwide, there are approximately 400 million traditional people liv-
ing in more than 70 countries, which occupy about 20% of the Earth’s land
surface (Scariot 2013; indigenouspeople.net). Rather than being a threat to
a pristine environment, in some cases traditional peoples have been an
integral part of these environments for thousands of years (Middleton
2013). The present mixture and relative densities of plants and animals
in many biological communities may reflect the historic activities—such
as fishing, selective hunting of game animals, and planting or encourag-
ing of useful plant species in fallow agricultural plots—of people in the
area. These activities often do not degrade the environment as long as
human population density remains low and there are abundant land and
resources.
Many traditional societies do have strong conservation ethics. These
ethics are often subtler and less clearly stated than Western conserva-
322 Chapter 9

tion beliefs, but they tend to influence people’s actions in their day-to-day
lives, perhaps more than Western beliefs (Ban et al. 2013). In such societies,
people use their traditional ecological knowledge to create management
practices that are linked to belief systems and enforced by village consent
and the authority of leaders. One well-documented example of such a
conservation perspective is that of the Tukano Indians in northwest Brazil
who have strong religious and cultural prohibitions against cutting the
forest along the Upper Río Negro, which they recognize as important to
the maintenance of fish populations (Andrew-Essien and Bisong 2009).
Local people who support conservation as an integral part of their
livelihood and traditional values are often inspired to take the lead in
protecting biodiversity. Empowering them by helping them to obtain legal
title —a right to ownership that is recognized by the government—to their
traditionally owned lands is often an important component of efforts to
establish locally managed protected areas in developing countries (Rai and
Bawa 2013). Today, indigenous communities own 97% of the land in Papua
New Guinea. Reserves for indigenous people in the Amazon Basin of Brazil
occupy over 100 million ha (22%) of its incredibly diverse habitats. The Inuit
people govern one-fifth of Canada. In Australia, tribal people control 90
million ha, including many of the most important areas for conservation.
Together these regions encompass a substantial percentage of the world’s
biodiversity (Figure 9.10).
The challenge, then, is to develop strategies for including these local
peoples in conservation programs and policy development both outside
and inside protected areas (Gavin et al. 2015). The partnership of tra-
ditional people, government agencies, and conservation organizations
working together has been termed co-management (Borrini-Feyerabend
et al. 2004). Co-management involves sharing management decisions and
their consequences. The new strategies have been developed in an effort
to avoid ecocolonialism, the practice by some governments and conserva-
tion organizations of disregarding the traditional rights and practices of
local people in order to establish new conservation areas. The practice
is called ecocolonialism because of its similarity to the historical abuses
of native rights by colonial powers of past eras (Cox and Elmqvist 1997).
The involvement of these people in the conservation of their lands is an
issue of social justice issue; this aspect will be discussed in more detail
in Chapter 11.
In many new conservation projects, the economic needs of local people
are included in conservation management plans, to the benefit of both the
people and the reserves (Roe et al. 2013). Such projects, known as integrated
conservation development projects (ICDPs), are now regarded as worthy
of serious consideration, though in practice they are often problematic to
implement, as described later in the chapter. There are many possible strat-
egies that could be classified as ICDPs, ranging from wildlife management
projects to ecotourism, and may or may not include formally protected
areas. These projects normally attempt to combine the protection of bio-
Conservation Outside Protected Areas 323

Amazon River

Manaus

Palmas

Deforestation hotspots
Deforested and cerrado
Forest
Land owned by
indigenous people 600 km
Strictly protected areas
Roads

Figure 9.10 Large blocks of indigenous lands are important in the overall
conservation strategy for the Brazilian Amazon. Many national parks and other
protected area have been established since 2002. Human activities of logging,
farming, and ranching over the past decade have created an “arc of deforesta-
tion.” Development and deforestation are also associated with the expanding
network of paved and unpaved roads. (From Soares-Filho et al. 2010.)

diversity and the customs of traditional societies with aspects of economic


development, including reducing poverty, creating jobs, improving health,
and ensuring food security. A large number of such programs have been
initiated over the last 25 years, and they have provided opportunities for
evaluation and improvement. Involving local people in ongoing monitor-
ing efforts may increase information and also help to determine how the
people themselves perceive the benefits and problems of the project (Bra-
schler 2009). The hope of such projects is that the local people will decide

Introduction to Conservation Biology 1E Primack/Sher


324 Chapter 9

that sustainable use of their local resources is more valuable than destruc-
tive use of those resources and that these people will become involved in
biodiversity conservation. The following are some examples of the types
of ICDPs currently in practice.

Biosphere reserves
In UNESCO’s World Network of Biosphere Reserves, traditional people
are allowed to use resources from designated buffer zones around strictly
protected core areas (see Figure 8.19). The program is a successful example
of the ICDP approach, at least in terms of its widespread adoption of as a
model of conservation; there are 621 Biosphere Reserves in 117 countries,
covering over 260 million ha. The Biosphere Reserve Program recognizes
the role of people in shaping the natural landscape as well as the need to
find ways in which people can sustainably use natural resources without
degrading the environment.
One instructive example of a biosphere re-
serve is the Kuna Yala Indigenous Reserve on
the northeast coast of Panama. In this protected
area, which comprises 60,000 ha of tropical forest
and coral islands, 50,000 Kuna people in 60 vil-
lages practice traditional medicine, fishing, ag-
riculture, and forestry (Figure 9.11). Scientists
from outside institutions carry out management
research, and in the process they train and hire
local people as guides and research assistants.
The Kuna local government attempts to control
the type and rate of economic development in
the reserve. However, a change appears to be
occurring among the Kuna: traditional conser-
vation beliefs are eroding in the face of outside
influences, often in tandem with the growing
tourism industry, and younger Kuna are begin-
ning to question the need to rigidly protect the
reserve (Posey and Balick 2006). Also, the Kuna
people have had difficulties establishing a stable
organization that can manage the reserve and
work with external conservation and donor
groups. Furthermore, rising sea levels and de-
clining marine resources are forcing village
leaders to consider other options for their fu-
ture. This example illustrates how empowering
traditional people is no guarantee that biodiver-
sity will be preserved. This is particularly true
when traditions change or disappear, economic
Figure 9.11 Kuna boys fishing. (© Alvaro pressures for exploitation increase, or programs
Leiva/AGE Fotostock.) are mismanaged. The challenge will be to de-
Conservation Outside Protected Areas 325

termine a way to integrate conservation into the cultural


evolution of Kuna society, which cannot—and from an ethi- ICDPs involve local people
cal standpoint, should not—be prevented. in sustainable activities
that combine biodiversity
In situ agricultural conservation
conservation and
The long-term health of modern agriculture depends on the economic development.
preservation of the genetic variability maintained in local
varieties of crops cultivated by traditional farmers (Bisht et
al. 2007). One innovative suggestion has been for an interna-
tional agricultural body, such as the Consultative Group on International
Agricultural Research, to subsidize villages as in situ (in-place) custodians
of traditional varieties of crop species (Brush 2004). Along these lines, in
China, a government program that involves paying farmers to interplant
high-quality traditional and high-yielding hybrid rice varieties maintains
the crop’s genetic variability (Zhu et al. 2003). Villages that participate in
such programs have an opportunity to maintain their culture in the face
of a rapidly changing world.
A different approach linking traditional agriculture and genetic con-
servation is being used in arid regions of the American Southwest, with
a focus on dryland crops with drought tolerance (www.nativeseeds.org).
A private organization, Native Seeds/SEARCH, collects the seeds of 1800
traditional crop cultivars for long-term preservation. The organization
also encourages a network of 4600 farmers and other members to grow
traditional crops, provides them with the seeds of traditional cultivars,
and buys their unsold production. The value of this and related genetic
conservation programs is being increasingly understood (Jarvis et al. 2016).
Countries have also established special reserves to conserve areas con-
taining wild relatives and ancient landraces of commercial crops (Barazani
et al. 2008). Species reserves protect the wild relatives of wheat, oats, and
barley in Israel and of citrus in India.

Extractive reserves
In many areas of the world, traditional people have extracted products from
natural communities for decades and even centuries. The use, sale, and
barter of these natural products are a major part of people’s livelihood. Un-
derstandably, local people are very concerned about retaining their rights to
continue collecting natural products from the surrounding countryside. In
areas where such collection represents an integral part of traditional society,
the establishment of a national park that excludes the traditional collection
of products will meet with as much resistance from the local community
as will a landgrab that involves the exploitation of the natural resources
and their conversion to other uses. A type of protected area known as an
extractive reserve may present a sustainable solution to this problem.
One such example is found in the Brazilian Amazon, where the govern-
ment is trying to address the legitimate demands of local citizens by establish-
ing extractive reserves from which settled people collect natural materials,
326 Chapter 9

such as medicinal plants, edible seeds, rubber, resins, and Brazil nuts, in ways
that minimize damage to the forest ecosystem (Duchelle et al. 2012) (Figure
9.12). These extractive reserves, which comprise about 3 million ha, guarantee
the ability of local people to continue their way of life and guard against the
possible conversion of the land to cattle ranching and farming. However,
populations of large animals in extractive reserves are often substantially
reduced by subsistence hunting by local people, and the density of Brazil nut
seedlings is reduced by the intense collection of mature nuts.
Many countries in East and southern Africa have started aggressively
applying community development and sustainable harvesting strategies
in their efforts to preserve wildlife populations. Governments are attempt-
ing to develop programs to generate income from trophy hunting and
wildlife tourism that can be operated at the village level and provide clear
benefits to local people (Naidoo et al. 2016). One example is the Com-
munity Based Natural Resource Management program, in which local
communities working with the government sell sport-hunting rights of
high-value trophy species, such as lions and elephants, to safari companies
(see Chapter 3). Revenue is also generated through operating tourist facili-
ties. To maintain the needed densities of wildlife, the village community
must work together with government officials to prevent illegal hunting
(see the chapter opening photo).
There has been vocal support from some conservation biologists for
selling hunting licences as a means of conserving species, especially when

Figure 9.12 Extractive reserves established in Brazil provide a reason to


maintain forests. The trunks of wild rubber trees are cut for their latex, which
flows down the grooves into the cup. Later the latex will be processed and used
to make natural rubber products. (Photograph © Edward Parker/Alamy.)
Conservation Outside Protected Areas 327

local people are involved (e.g., Di Minin et al. 2016). However, trophy hunt-
ing is considered ethically questionable both those who believe that killing
purely for sport (rather than for food) is morally wrong and point to the
faulty reasoning behind consequentialism, that is, that the ends justify the
means (Nelson et al. 2016). Furthermore, they argue that revenues from
trophy hunting for conservation are insufficient, usually do not reach the
local community, and are decreased via corruption (Lindsey et al. 2016).
Finally, the market for hunting licenses creates pressure to “produce an ani-
mal” that inevitably leads to poaching, as was seen in the Cecil story (see
Chapter 3; Richard Reading, pers. comm). Thus, some question whether
sport hunting belongs in the same category as extractive reserves that
support locals with food, firewood, or other resources.

Community-based initiatives
In many cases, local people already protect natural areas and resources
such as forests, wildlife, rivers, and coastal waters in the vicinity of their
homes. Protection of such areas, sometimes called community conserved
areas, or community-based conservation (CBC), is often enforced by vil-
lage elders because of the clear benefit to the local people (see Chapter 11).
These benefits include maintaining natural resources (e.g., food supplies
and drinking water) and the use of the land for religious and traditional
practices. The protection of biodiversity may even be an intrinsic aspect
of local beliefs (Borrini-Feyerabend et al. 2004). In this way, the goal of a
CBC is to align ecological, economic, and social goals. A review of 136 CBC
projects across the globe found that degree of local participation, environ-
mental education and skills-training programs all significantly contributed
to win–win outcomes for the people and to biodiversity (Brooks 2016). The
most important feature, however, was institutional capacity building: ef-
forts to improve infrastructure and communication and decision-making
processes. Governments and conservation organizations can assist local
conservation initiatives by providing access to scientific expertise, train-
ing programs, and financial assistance to develop needed infrastructure,
in addition to simply offering legal title to traditional lands.
One example of a local initiative is the Community Baboon Sanctuary in
eastern Belize, which was created by a collective agreement among a group
of villages to maintain the forest habitat required by the local population of
black howler monkeys (known locally as baboons). Ecotourists visiting the
sanctuary pay a fee to the village organization, and additional payments
are made if they stay overnight and eat meals with a local family. Conserva-
tion biologists working at the site have provided training for local nature
guides, a body of scientific information on the local wildlife, funds for a
local natural history museum, and business training for the village leaders.
In the Pacific islands of Samoa, much of the rain forest land and marine
area is under “customary ownership”: it is owned by communities of indig-
enous people (Boydell and Holzknecht 2003). Villagers are under increasing
pressure to sell logs from their forests to pay for schools and other necessi-
328 Chapter 9

ties. Despite this situation, the local people have a strong desire to preserve
the land because of the forest’s religious and cultural significance, as well
as its value for medicinal plants and other products. A variety of solutions
are being developed to meet these conflicting needs. In 1988, in American
(or Eastern) Samoa, where about 90% of the land is under customary owner-
ship, the US government leased forest and coastal land from the villages
to establish a new national park (americansamoa.noaa.gov). Under this
agreement, the villages gained needed income yet retained ownership of
the land and their traditional hunting and collecting rights (www.nps.gov).

Payments for ecosystem services


A creative strategy involves making direct payments to individual land-
owners and local communities that protect critical ecosystems and the
services they provide, in effect paying the community to be a good stew-
ard of the land (Wunder 2013; Wünscher and Engel 2012). These types of
programs are sometimes referred to as payments for ecosystem services
(PES), and they are becoming increasingly popular (Figure 9.13). Govern-
ments, nongovernmental conservation orga-
nizations, and businesses develop markets
(A)
in which local villagers and landowners can
100 participate through protecting and restor-
Frequency (number of projects)

ing ecosystems. For example, owners of a


80
forest may receive direct payments from a
60
city government for the ecosystem servic-
es provided by the forest, such as control-
40 ling floods and providing drinking water.
Local landowners and farmers can be paid
20 for allowing large predators such as wolves,
bears, tigers, and mountain lions to be on
0 their land, with additional payments given
Habitat Over- Pollution Climate Invasives as compensation if their livestock is attacked
conversion harvesting change
(Dickman et al. 2011).
Threat type
Rural people can be drawn into newly
(B) developing international markets for eco-
120

100 Figure 9.13 Patterns of payments for 103


Number of funders

ecosystem services (PES) projects from 37


80
countries. (A) Number of projects addressing
60 different types of threat. Most projects address
issues of habitat conversion (from forest to ag-
40 ricultural land) and overharvesting of trees. (B)
Funding sources for the projects are primarily
20
nonprofit (NP) conservation organizations but
0 also include government agencies at national,
NP Federal Corporations State Local state, and local levels, as well as corporations.
Funder type (After Tallis et al. 2009.)
Conservation Outside Protected Areas 329

system services (www.ecosystemmarketplace.com), especially


programs that trade in carbon credits for reducing atmospheric New markets are being
levels of greenhouse gases. In the Kasigu Corridor Reducing developed in which local
Emissions from Deforestation and Forest Degradation (REDD) people and landowners
project in southeastern Kenya, villagers protect a natural migra- are paid for providing
tion corridor for elephants between Tsavo East National Park ecosystem services, such
and Tsavo West National Park (Figure 9.14). The villagers earn as protecting forests to
carbon credits and funding from international programs for
maintain water supplies
protecting the forest and maintaining wildlife populations, as
and planting trees to
determined by an independent outside evaluation. Funds from
absorb carbon dioxide.
the program have been used to pay for wildlife patrols, build
Programs that address
schools and other infrastructure, and start local businesses,
while in the process creating 350 new jobs and generating a climate change issues are
total of $1.2 million (Dinerstein et al. 2012). Programs addressing predicted to become more
carbon sequestration and climate change are likely to expand common in the coming
greatly in coming years, and they may provide substantial funds years.
for land protection. However, at present such programs are
sometimes unable to pay enough money to prevent landowners
from converting their land to other uses (Banerjee et al. 2013).
PES have been effective for improving habitat for the giant panda (Ai-
luropoda melanoleuca) in China, as a part of the Natural Forest Conservation
Program (Tuanmu et al. 2016). In fact, an evaluation of the changes in veg-
etation revealed that often more improvement occurred in areas managed
by locals than those managed by the government. However, this success
only took place when the financial incentives were very high.

Figure 9.14 Elephants (Loxodonta africana) in Kenya depend on this wildlife


corridor for their annual migration between two national parks. The corridor is
maintained by locals who benefit financially. (© Morkel Erasmus/Getty Images.)
330 Chapter 9

Evaluating conservation initiatives that involve traditional


societies
Unfortunately, when external funding ends, and if the projected income
stream fails to develop, many of these integrated conservation and devel-
opment projects (ICDPs) end abruptly. Even for projects that appear suc-
cessful, there is often no monitoring of ecological and social parameters
to determine whether the project goals are being achieved. It is essential
that any conservation program design include mechanisms for evaluating
the progress and success of measures being taken.
A key element in the success of many of the projects discussed in the
preceding sections is the opportunity for conservation biologists to com-
plement and work with stable, flexible, local communities with effective
leaders and competent government agencies (Baker et al 2012). Certain
projects appear to be successful at combining the protection of biodiversity
with sustainable development and poverty reduction. The Equator Initia-
tive of the United Nations is cosponsored by many leading conservation
organizations, businesses, and governments and is helping to fund and
publicize such efforts.
However, in many cases a local community may have internal conflicts
and poor leadership, making it incapable of administering a successful
conservation program. Moreover, conservation initiatives involving recent
immigrants or impoverished, disorganized local people may be difficult to
carry out, and government agencies working on the project may be ineffec-
tive or even corrupt. Consequently, while working with local people may
be a desirable goal, in some cases it simply is not possible.

Case Studies: Namibia and Kenya


Throughout the world, the protection of biodiversity is being included as
an important objective of land management. We conclude the chapter by
examining two case studies of successful community-based programs in
Namibia and Kenya that illustrate some of the challenges and successes
of managing biodiversity outside protected areas.
Community-Based Natural Resource Management (CBNRM) programs
in Africa represent an approach in which local landowners and communal
groups are given the authority to manage and profit from the wildlife on
their own property. In many African countries, wildlife both inside and
outside national parks is often managed by government officials, often
with no input from the local people, who gain little or no economic benefit
from the wildlife on their own land and have no incentive to protect the
wildlife. By changing the management system to CBNRM, government
officials and conservation organizations hope to counterbalance pressures
threatening local wildlife while simultaneously contributing to rural eco-
nomic development. There is a long history of CBNRM programs in Africa,
but it has been difficult to develop stable programs that are economically
viable and effectively managed.
Conservation Outside Protected Areas 331

One of the most ambitious new programs for local communities man-
aging wildlife is found in Namibia in southern Africa (Riehl et al 2015).
Namibia has over 2.3 million people, with 62% of them living in rural
areas and farming or raising livestock (NACSO 2014). Namibia includes
an impressive six different biomes, from the desert to the subtropical
(Figure 9.15). These biomes support high levels of endemism, includ-
ing 700 plant species, 91 bird species, and 26 mammal species (UNCBD
report 2010).

Angola

Nyae Nyae
community conservancy

Windhoek Botswana
Atlantic Ocean Namibia
Community conservancies
Other conservancies
Protected areas
Biome
Namib Desert
Semi-desert
Lakes and salt pans
Tree and shrub savanna

250 km
South Africa 150 miles

Figure 9.15 The distribution of current and emerging community conser-


vancies in Namibia, in which communal groups agree to protect biodiversity.
State-protected lands are also shown. The “Other conservancies” category de-
notes areas in which communal groups have not yet committed to forming con-
servancies to protect wildlife. It can be seen here that several biomes are more
represented in these community conservancies than in formal protected areas
(After NACSO 2008, with updates from Riehl et al. 2015.)
332 Chapter 9

Beginning in 1996, the Namibian government granted traditional com-


munal groups the right to use and manage the wildlife on their own lands.
To obtain these rights, a group needs to form a management committee
and determine the boundaries of its land. The government then designates
the group as a “community conservancy.” The benefits of forming a con-
servancy and participating in wildlife management are fourfold:
1. The conservancy can form joint ventures with tour operators, with
about 5–10% of the gross earnings paid to the conservancy. A certain
number of the employees in the tourist operation are hired from among
the communal group. Revenues from the joint ventures are used to train
and pay game guards, again hired from the communal group, who
monitor the wildlife populations and prevent poaching.
2. Using funds from the joint ventures, the conservancy members can
build and operate campsites for tourist groups, providing direct rev-
enue, employment, and experience for the communal group.
3. The conservancy can apply to the government for a trophy-hunting
quota, which will be granted if the wildlife populations are large
enough, as indicated by monitoring. Hunting licenses can then be sold
or auctioned off to professional hunters, who bring in wealthy foreign
tourists willing to pay a high price for an African hunting experience.
Payments to the conservancy for high-value animals such as lions and
elephants can be as large as $11,000 per animal. Meat from the hunted
animals is distributed to the group members as an added benefit. This
approach to funding is not without controversy, however (see Chapter
3 and “Extractive reserves” in this chapter). Some economic analysis
suggests that tourism alone (i.e., without hunting) is not sufficient to
cover operating costs (Naidoo et al. 2016).
4. Once the conservancy has formed a wildlife management plan, four
species of wildlife—gemsbok, springbok, kudu, and warthog—can be
hunted for subsistence. In practice, the hunting is often done by game
guards and professional hunters and the meat is distributed to every-
one in the community.

Over the last 18 years, 79 conservancies in Namibia have been estab-


lished, covering 19.4% of Namibia’s land surface (Riehl et al. 2015) (see Figure
9.15). Help in the initial establishment of the conservancies has come from
external funding agencies, such as the US Agency for International Devel-
opment. Conservancy members have received further training in tourism,
finance, and marketing, along with effective advocacy to gain support from
the government and the private sector. Although the financial gain by the
CBNRM is well documented and social benefits such as improved health
in these areas relative to adjacent areas have been documented, whether
other expected social benefits, such as education, have improved is less
clear (Riehl et al. 2015). This is due in part to the large degree of variability
Conservation Outside Protected Areas 333

between conservancies with a great deal of tourism and those that have
none. Some analyses also suggest that although hunting and tourism in
conservancies could earn more per hectare than livestock rearing (Lindsey
et al. 2013), conservancies may not be economically viable in the long term
(Humavindu and Stage 2015). PES and other revenue-sharing systems have
been proposed to address the problem of viability (Lapeyre 2015).
So far, the communal management system seems to be having posi-
tive effects on conservation. Namibia currently claims to host the world’s
largest populations of free-ranging cheetah and black rhino, both inter-
national species of concern (see chapter opening photo). A report from
a consortium of conservation nongovernmental organizations (NGOs),
including WWF Namibia, working with the Ministry of Environment,
documented dramatic increases in many large mammals since the
CBNRM programs were initiated (NASCO 2008). Many species, especially
ungulates, have been observed as having greater diversity and higher
numbers within the conservancies than in adjacent, unprotected land
(Lindsey et al. 2015).
Other African countries have programs that are similar to Namibia’s.
In Kenya, for example, about two-thirds of the country’s 650,000 large
animals—including giraffes, elephants, zebras, and ostriches—live outside
park boundaries in rangelands used by commercial ranches and as tradi-
tional grazing lands by local people (Western et al. 2009; Young et al. 2005).
The rangelands outside the parks are increasingly unavailable to wildlife,
though, because of fences, poaching, and agricultural development, which
have led to a gradual decline in wildlife numbers.
A combination of regulations, community involvement, and economic
incentives is contributing to the persistence of substantial populations of
wildlife in certain of these unprotected areas of Kenya in spite of the chal-
lenges (Kinnaird and O’Brien 2013). In some places, private ranching in
which wildlife and livestock are managed together for both meat and
ecotourism is more profitable than managing livestock alone because the
livestock and the wildlife use different food resources. As in Namibia,
many ranches have also developed facilities for foreign tourists who want
to view wildlife, which creates an additional source of revenue and an
incentive for protecting these species.
Although these community-based management programs have been
successful in many cases, their dependence on tourism and subsidies from
outside donor governments and conservation and development organiza-
tions can make them vulnerable. When these outside subsidies cease, the
wildlife programs often end as well, suggesting that the programs are
often not really profitable on their own. The ineffectiveness and corruption
of some local government agencies are additional factors that can cause
such programs to fail. These community wildlife programs will be judged
successful when they can demonstrate that they can both protect wildlife
and provide a stable income source for the local people.
334 Chapter 9

Summary
„„Considerable biodiversity exists outside „„Government agencies, private conser-
protected areas, particularly in habitat vation organizations, businesses, and
managed for multiple-use resource ex- private landowners can cooperate in
traction. Such unprotected habitats are large-scale ecosystem management proj-
vital for conservation because in almost ects to achieve conservation objectives
all countries, protected areas account and use natural resources sustainably.
for only a small percentage of total area. Bioregional management involves coop-
Animals and plants living in protected eration across large regions to manage
areas often disperse to unprotected land, large ecosystems, which frequently cross
where they are vulnerable to hunting/ international borders.
harvest, habitat loss, and other threats „„In Africa, many of the characteristic
from humans. large animals are found predominantly
„„Governments are increasingly encour- in rangeland outside the parks. Local
aging the protection of biodiversity as people and landowners often maintain
a priority on multiple-use land, includ- wildlife on their land for a variety of
ing forests, grazing lands, agricultural purposes. Local communities are now
areas, military reservations, and urban generating income by combining wild-
areas. All of these can be managed for life management and ecotourism, some-
conservation, keeping in mind that there times including trophy hunting.
are species that are too sensitive to ever
exist outside of strictly protected areas.

For Discussion
1. Consider a national forest that has been what basis should this decision be made:
used for decades for logging, hunting, economic, ethical, past success, or future
and mining. If endangered plant species potential for conservation?
are discovered in this forest, should these 3. Choose a large aquatic ecosystem that
activities be stopped? Can logging, hunt- includes more than one country, such as
ing, and mining coexist with endangered the Black Sea, the Rhine River, the Carib-
species, and if so, how? If logging has to bean, the St. Lawrence River, or the South
be stopped or scaled back, do the logging China Sea. What agencies or organiza-
companies or their employees deserve tions have responsibility for ensuring the
any compensation? Explain your answer. long-term health of the ecosystem? In
2. Do you think that trophy hunting on pri- what ways do they, or could they, cooper-
vate reserves is a good means by which ate in managing the area?
to preserve species? Why or why not? On
Conservation Outside Protected Areas 335

Suggested Readings
Athreya, V., M. Odden, J. D. Linnell, J. Krishnaswamy, and U. Karanth. 2013. Big
cats in our backyards: Persistence of large carnivores in a human dominated
landscape in India. PLoS ONE 8(3): e57872. Leopards live near farmlands
and villages but do not affect people or livestock because they hunt at night.
Baudron, F. and K. E. Giller. 2014. Agriculture and nature: Trouble and strife?
Biological Conservation 170: 232–245. The authors consider the alternatives
of land sharing and land sparing.
Brooks, J. S. 2016. Design features and project age contribute to joint success in
social, ecological, and economic outcomes of community-based conserva-
tion projects. Conservation Letters. Capacity building, particularly when it
improves infrastructure, is the most important predictor of success in CBC
projects.
Davis, A. M. and A. R. Moore. 2015. Conservation potential of artificial water
bodies for fish communities on a heavily modified agricultural floodplain.
Aquatic Conservation: Marine and Freshwater Ecosystems. Many species of
fish can be found in gravel pits and other haphazard water bodies, but the
greatest diversity was found in constructed wetlands.
Farmer, J. R., Z. Ma, M. Drescher, E. G. Knackmuhs, and S. L. Dickinson. 2016.
Private landowners, voluntary conservation programs, and implementa-
tion of conservation friendly land management practices. Conservation Let-
ters. Attitudes and motivations of private landowners affect how well they
manage for biodiversity.
Gavin, M. C., J. McCarter, A. Mead, F. Berkes, J. R. Stepp, D. Peterson, and
R. Tang. 2015. Defining biocultural approaches to conservation. Trends in
Ecology and Evolution 30(3): 140–145. For conservation to work, local people
must be involved.
Hylander, K., S. Nemomissa, J. Delrue, and W. Enkosa. 2013. Effects of coffee
management on deforestation rates and forest integrity. Conservation Biol-
ogy 27: 1011–1019. Traditional forms of coffee plantations can maintain for-
est cover and some level of biodiversity.
Lindsey, P. A., C. P. Havemann, R. M. Lines, A. E. Price, T. A. Retief, T. Rhe-
bergen, and 2 others. 2013. Benefits of wildlife-based land uses on private
lands in Namibia and limitations affecting their development. Oryx 47(1):
41–53. CBNRM is effective for protecting large mammals, especially ungu-
lates, in part due to their management for sport hunting.
Loke, L. H., R. J. Ladle, T. J. Bouma, and P. A. Todd. 2015. Creating complex
habitats for restoration and reconciliation. Ecological Engineering, 77: 307–
313. Sometimes that which benefits people can also benefit biodiversity.
Naidoo, R., L. C. Weaver, R. W. Diggle, G. Matongo, G. Stuart-Hill, and C.
Thouless. 2016. Complementary benefits of tourism and hunting to com-
munal conservancies in Namibia. Conservation Biology. Tourism and hunt-
ing are distinct funding sources and become profitable at different periods
during a conservancy’s development.
Riehl, B., H. Zerriffi, and R. Naidoo. 2015. Effects of community-based natural
resource management on household welfare in Namibia. PLoS ONE 10(5).
The benefits of this approach can be seen on multiple levels of society but
not always in ways one would expect.
10 Restoration Ecology
Where to Start? 339 Restoration of Some Major
Restoration in Urban Areas 344 Communities 351
Restoration Using Organisms 346 The Future of Restoration
Ecology 359
Moving Targets of Restoration 350

Volunteers and fishermen plant mangrove trees in Kondang Merak, located


on the eastern coast of Java, Indonesia, in an effort to restore the local
marine ecosystem. This area has been heavily damaged by illegal fishing
methods that have included the use of cyanide and dynamite.
E
cosystems that have been damaged or de-
stroyed by intensive human activities such
as mining, ranching, and logging may lose
much of their ecological resilience, or natural ability
to recover. In some cases, recovery would require cen-
turies or even millennia without human assistance.
The restoration of ecosystems can be motivated by
the protection of species, improving aesthetics, rec-
reation, strengthening ecosystem function or connec-
tivity, or the reestablishment of ecosystem services
(Keenelyside 2012). Rebuilding damaged ecosystems
also can be used to enlarge, enhance, and connect
protected areas, as well as to create buffers around
them (see Chapter 8).
Degraded ecosystems provide important opportu-
nities to apply research findings by helping to restore
historical species and communities (Clewell and Aron-
son 2008). Ecological restoration is the practice of
restoring the species and ecosystems that occupied
a site at some point in the past but were damaged or
destroyed (www.ser.org) (Figure 10.1). Restoration
ecology is the science of restoration—the research and
scientific study of restored populations, communities,
338 Chapter 10

Figure 10.1 (A) Trout stream habi- (A)


tat that has been degraded by human
activities. (B) Trout stream habitat that
has been restored by installing fencing
to exclude cattle, planting native species,
and reinforcing stream banks with rocks.
(From Hobbs et al. 2010; photographs
courtesy of K. Matthews.)

(B)

and ecosystems (Falk et al. 2006). These are overlapping disci-


Some ecosystems have plines: the process of ecological restoration provides useful sci-
been so degraded by entific data, while restoration ecology interprets and evaluates
human activity that their restoration projects in a way that can lead to improved methods.
resilience, or ability to Restoration ecology will play an increasingly important
recover on their own, role in the conservation of biodiversity as degraded lands and
is severely limited.
aquatic communities are partially or completely restored to
their original species compositions and are integrated into
Ecological restoration
existing conservation reserve networks. Because many de-
reestablishes functioning
graded areas are unproductive and of little economic value,
ecosystems, with some or
governments may be willing to restore them to increase their
all of the original species
economic productivity and conservation value. For example,
or, sometimes, a different degraded areas may be subject to soil erosion and increased
group of species. risk of flooding; restoration in these cases may be motivated by
a desire to mitigate threats to human life or property. Restora-

Introduction to Conservation Biology 1E Primack/Sher


Sinauer Associates
Morales Studio/SA
Restoration Ecology 339

tion efforts also can be part of compensatory mitigation or biodiversity


offsets, in which a new site is created or rehabilitated in compensation
for a site that has been destroyed elsewhere by development (Maron et
al. 2012). This is particularly true for wetlands, for which a “no net loss
policy” has been adopted by many jurisdictions. At other times, ecologi-
cal processes rather than ecosystems need to be restored. For example, an-
nual floods disrupted by the construction of dams and levees or natural
fires stopped by efforts at fire suppression may need to be reintroduced if
the absence of these processes proves harmful to species and ecosystems.
Ecological restoration has its origins in older, applied technologies that
attempted to restore ecosystem functions or species of known economic
value, such as wetland creation (to prevent flooding), mine site reclama-
tion (the final stage of closing a mine, usually involving plantings to
prevent soil erosion), range management of overgrazed lands (to increase
the production of grasses), and technologies to facilitate tree planting on
cleared land (for timber, recreational, and ecosystem values). However,
these approaches often produce biological communities that are overly
simplified or cannot maintain themselves. As concern for biodiversity
has grown, restoration plans have included as a major goal the rees-
tablishment of original or historical species assemblages and processes.
For example, one of the objectives of the European Union Biodiversity
Convention was to restore at least 15% of degraded ecosystems by 2020
(European Commission 2011). The input of conservation biologists is
needed to achieve these goals.

Where to Start?
To be successful in the long term, restoration projects must first establish
clear goals, followed by an assessment of site conditions to determine if
those goals can be met, and if so, by what means (Figure 10.2). Often,
the goal of restoration efforts is to create ecosystems that are comparable
in function or species composition to existing reference sites (Humphries
and Winemiller 2009). Reference sites are central to the very concept of res-
toration; they act as comparison sites, providing explicit restoration goals
and allowing for quantitative measures of the project’s success (Higgs et
al. 2014). Unrestored areas can act as “negative” reference sites or controls
to further determine the impact of the restoration actions.
If practical, the successfully restored ecosystem should be dominated
by native species, contain representatives of all key functional groups of
species, have a physical environment suitable for native species and eco-
system processes, and be secure from detrimental outside disturbances.
In some cases, such as at arid and cold sites, achieving such recovery may
take decades or even centuries.
But is the reestablishment of native species always the goal of restora-
tion? Site conditions or limitations of resources may make this undesirable
or impossible. There are four main approaches that define outcomes when
340 Chapter 10

A. Determine overarching restoration goal

B. Evaluate non- D. Establish a realistic C. Evaluate ecological factors:


ecological factors: restoration objective: • Site history
• Site logistics • Given available resources • Existing physical and
• Relevant policies and constraints, what is the biotic conditions
• Financial resources desired end-state that would • Ecological processes
• Human resources accomplish the goal? (e.g., flood, fire, erosion)

E. Create site-specific restoration plan:


• How will project objectives be accomplished?
• How will pre- and post-restoration monitoring be done?

F. Conduct pre-project monitoring

G. Implement plan

H. Conduct post-project monitoring

I. Engage adaptive management Link to related projects

Figure 10.2 A flow diagram of a scientific approach to restoration. The first


step is to (A) clarify the overarching goal of the restoration project, such as to
increase biodiversity at a site. Then, to determine how this will be achieved, it is
necessary to evaluate both (B) nonecological factors, such as how much money
is available for the project, and (C) ecological factors, such as the condition of the
soil and what plant species currently are found there. These factors will inform
and be influenced by (D) the specific objectives for restoration, such as estab-
lishing plant species that will promote diversity at higher levels. Only after these
steps have been taken can (E) a specific, realistic plan for how to implement
restoration be created, such as removing weeds, improving the soil, or planting
seeds. The implementation of the plan should accompany both pre- and post-
monitoring, (F–H). Monitoring progress of the restoration objectives will then
facilitate (I) improvement over time at both that site and future projects. (After
Shafroth et al. 2008.)

considering the restoration of biological communities and ecosystems


(Figure 10.3) (Bradshaw 1990):
Introduction to Conservation Biology 1E Primack/Sher
1. No action. Restoration is deemed too expensive, previous attempts have
Sinauer Associates
Morales Studiofailed, or experience has shown that the ecosystem will recover on its
Primack_Sher1E_10.02 Date 02-24-16 3-3-16 3-7-16 3-15-16
own. Letting the ecosystem recover on its own, also known as passive
restoration, is typical for old agricultural fields in eastern North Amer-
Restoration Ecology 341

Rehabilitation:
High Rehabilitation:
Replacement of
ORIGINAL
many species,
Replacement of new ecosystem ECOSYSTEM
a few species
Complete restoration
Biomass, nutrient content, etc.
ECOSYSTEM FUNCTION

Partial restoration

No action: Ecosystem recovers


on its own via succession

DEGRADED ECOSYSTEM
No action: Continued
Low

deterioration
Low Number of species and ecosystem complexity High
ECOSYSTEM STRUCTURE

Figure 10.3 Decisions must be made about whether the best course of
action is to completely restore a degraded site (green arrow), partially restore it
(blue arrow), rehabilitate it by introducing different species (black arrow), or take
no action (red arrow). (After Bradshaw 1990.)

ica, which sometimes return to forest within a few decades after being
abandoned. However it should be noted that even in this case, the spe-
cies composition may be quite different, especially for the understory
(Flinn and Marks 2007).
2. Rehabilitation. A degraded ecosystem is replaced with a different but
productive ecosystem type. For example, a degraded forest might be
replaced with a productive pasture or a tree plantation. Just a few spe-
cies may be replaced, or a larger-scale replacement of many species
may be attempted. As the ultimate goal is not to restore the original
ecosystem, some authors consider the term restoration inappropriate to
refer to rehabilitated ecosystems (Perring et al. 2014).
3. Partial restoration. At least some of the ecosystem functions and some of
the original, dominant species are restored. An example is replanting a
degraded grassland with a few species that can survive. Partial restora-
tion typically focuses on dominant species or particularly resilient species
that are critical to ecosystem function, delaying action on the rare and less
common species that would be part of a complete restoration program.
4. Complete restoration. The area is completely restored to its original spe-
cies composition and structure by an active program of site modifica-
tion and reintroduction of the original species. For complete restoration,
the first step is to determine and then mitigate the source of ecological
degradation. For example, the source of pollution of a lake ecosystem

Introduction to Conservation Biology 1E Primack/Sher


Sinauer Associates
342 Chapter 10

must be identified and controlled before the ecosystem can be restored.


Natural ecological processes must be reestablished because they help
the system recover and contribute to long-term resilience.
In practice, ecological restoration must also consider the speed of res-
toration, the cost, the reliability of results, and the ability of the target
community to persist with little or no further maintenance. Practitioners
of ecological restoration must have a clear grasp of how natural systems
work and what methods of restoration are feasible (Falk et al. 2006). Con-
siderations of the cost and availability of seeds, when to water plants,
how much fertilizer to add, how to remove invasive species,
and how to prepare the surface soil may become paramount in
Restoration projects determining a project’s outcome. Dealing with such practical
require monitoring to details generally has not been the focus of academic biologists
determine whether goals in the past, but they must be considered in ecological restora-
such as costs and speed tion. Fortunately, restoration ecology often involves profession-
of recovery are being met. als from other fields, who can lend their expertise, ultimately
Such projects may also enriching the restoration process. However, these practitioners
provide new insights into sometimes have different goals than conservation biologists.
ecological processes.
For instance, civil engineers involved in major projects seek
economical ways to permanently stabilize land surfaces, prevent
soil erosion, make the site look better to the general public, and
if possible, restore the productive value of the land.
The degree of alteration will generally dictate what type of action is
required; the most degraded systems will likely require both biological
and physical alterations to the habitat (Figure 10.4). If the damage has
been caused by abiotic factors such as soil erosion or lowered water tables,
then the source of the problem should be addressed or at least considered
before making any attempt to reestablish species (Hobbs and Harris 2001).
One particular quandary in the restoration of highly degraded ecosys-
tems is ecological assembly order; that is, in what order and when should
the species components of the system be put back together? When trophic
order is apparent (e.g., predators require an established prey species), the
decision seems obvious, but the overlap in functional relationships among
many species complicates such decisions. For example, degraded parks in
Africa may have contained a dozen or more ungulate species with similar
ecological roles. In what order should these be reintroduced? Knowledge
about the specific ecology of individual species can help guide such com-
plicated decisions (Temperton and Hobbs 2004).
Another important issue is the genotypes of plant species that are being
reintroduced to a restoration site. In general, it is advisable to follow the
local-is-best (LIB) approach, which prioritizes locally adapted genotypes.
This is important for at least two reasons. First, there may be local adapta-
tions within a species that will make it either more or less likely to thrive
at the restoration sites. Second, there is a risk of outbreeding depression
or genetic swamping of local strains by introducing non-local genotypes
(see Chapter 5). There are also problems with this approach, including
Restoration Ecology 343

Fully
functional Requires Requires Requires

Biotic barrier
Abiotic barrier
physical- biological improved
chemical modification management
modification
Ecosystem attribute

Non-
functional
Degraded Intact
Ecosystem state

Figure 10.4 A conceptual model that considers thresholds for the restora-
tion of ecosystem function. Generally, the most degraded, and therefore non-
functional, sites will require overcoming abiotic constraints that contribute to
the problem, such as removing levees from a river, building structures for coral
to grow on, or amending soil that is too acidic. Biotic barriers can be overcome
by planting, reintroducing missing trophic levels, or providing a food source.
Of course, biotic changes can lead to abiotic ones, as in the case of an invasive
plant species increasing the frequency or intensity of fires. If an ecosystem is
mostly intact and functional, improved management rather than restoration is
needed. (Parks Canada and the Canadian Parks Council, 2008; after Whisenant
1999, and Hobbs and Harris 2001.)

that such genetic matching may be unfeasible or impossible (Smith et al.


2007). In other cases, harvesting wild native plants risks hurting the source
populations (Meissen et al. 2015), or the local genotypes have such low
genetic variability as to risk inbreeding depression at the restoration site.
For this reason, restoration practitioners may need to resort to less related
plant stock or even nonnative species (Jones 2003). In at least some cases, the
use of cheaper, commercial seed rather than seed that was hand collected
or specially reared has no adverse ecological effects (Reiker et al. 2015).
Introduction to Conservation Biology 1E Primack/Sher
Once
Sinauer restoration has begun, to determine whether these goals are being
Associates
achieved,
Morales Studioboth the restoration and the reference sites must be monitored
Primack_Sher1E_10.04
over time (González Date
et 12-18-15
al. 2015). Because restoration efforts need to be
customized for individual sites, it is often advisable to experimentally test
different restoration methods (Lloyd et al. 2013). These restored sites then
need to be monitored for years or even decades to determine how well
management goals are being achieved and whether further intervention
is required, an approach called adaptive management or adaptive resto-
344 Chapter 10

ration (Wagner et al. 2008). In particular, native species may have to be


reintroduced if they have not survived, and invasive species may have to
be removed if they are still abundant or there has been a secondary inva-
sion by a different invader.
Restoration ecology is valuable to the broader science of ecology be-
cause it provides an acid test of how well we understand a biological com-
munity; the extent to which we can successfully reassemble a functioning
ecosystem from its component parts demonstrates the depth of our knowl-
edge and points out deficiencies (Bradshaw 1990). For example, Grman and
colleagues (2014) defined rules for the assembly of plant communities on
prairies after examining 29 restoration projects in southwestern Michigan.
Efforts to restore degraded terrestrial communities generally have em-
phasized the establishment of the original plant community, as it typically
contains the majority of the biomass and provides structure for animals
and other elements of the community. However, some researchers argue
that in the future, restoration ecology should devote more attention to the
other major components of the community (Fraser et al. 2015). Fungi and
bacteria (see Chapter 3) play vital roles in soil decomposition and nutrient
cycling; soil invertebrates are important in creating soil structure; herbivo-
rous animals are important in reducing plant competition and maintaining
species diversity; birds and insects are essential pollinators; and many
birds and mammals have vital functions as insect predators, soil diggers,
and seed dispersers (Morandin and Kremen 2013). Many birds, insects, and
other animals may be able to recolonize the site on their own, but other
large animals and above-ground invertebrates may have to be reintroduced
from existing populations or captive breeding populations (see Chapter 7)
if they are unable to disperse to the site on their own (Audino et al. 2014).
Restoration efforts may also succeed by focusing on reestablishing ecologi-
cal processes that support native communities rather than just planting
specific plant taxa (Moreno-Mateos et al. 2015b).
Another practical consideration in restoration projects is gaining the
support and participation of local stakeholders. Many restoration efforts
are supported and even initiated by local conservation groups because
they recognize the direct connection between a healthy environment and
people’s personal and economic well-being (Felson et al. 2013; Higgs 2003)
(see Chapter 3).

Restoration in Urban Areas


Highly visible restoration efforts are taking place in many urban areas.
These efforts seek to reduce the intense human impact on ecosystems and
enhance the quality of life for city dwellers (Felson et al. 2013; Shwartz et
al. 2014). Local citizen groups often welcome the opportunity to work with
government agencies and conservation groups to restore degraded urban
areas. Unattractive drainage canals in concrete culverts can be replaced
with winding streams bordered with large rocks and planted with native
Restoration Ecology 345

Figure 10.5 The Cheonggyecheon Stream in the center of Seoul, Korea, is an


example of an urban restoration project that has provided many ecosystem ser-
vices and improved the quality of life for residents. (© Alex Barlow/Getty Images.)

wetland species. Vacant lots and neglected lands can be replanted with
native shrubs, trees, and wildflowers. Gravel pits can be packed with soil
and restored as ponds. Establishing native plant species in these urban
areas often leads to increases in populations of native birds and insects
(Burghardt et al. 2009). These efforts have the additional benefits of foster-
ing neighborhood pride, creating a sense of community, and enhancing
property value (Figure 10.5). However, such restorations are often only
partially successful because of their small size and the fact that they are
embedded in the highly modified urban environment. Developing urban
places where people and biodiversity can coexist has been termed recon-
ciliation ecology (Rosenzweig 2003).
An example of the value of restoration projects to people in urban set-
tings is evident in Japan, where parents, teachers, and children in Tokyo
and Yokohama have built over 500 small ponds next to schools
and in public parks to provide habitat for dragonflies and other
native aquatic species (Kobori 2009). The ponds are planted with
aquatic plants; many dragonflies colonize them on their own, Highly visible restoration
and some species are carried in as nymphs from other ponds. efforts are taking place
Dragonflies are an important symbol in Japanese culture, and in many urban areas to
dragonfly ponds are useful for teaching zoology, ecology, chem- reduce the intense human
istry, and principles of conservation. The schoolchildren are impact on ecosystems
responsible for the regular weeding and maintenance of these and enhance the quality
“living laboratories,” which helps them to feel an ownership of of life for city dwellers.
the project and to develop environmental awareness.
346 Chapter 10

Restoring native communities on huge landfills presents one of the


most unusual opportunities for urban restoration. In the United States,
150 million tons of trash are buried in over 5000 active landfills each year.
When the landfills have reached their maximum capacity, they are usually
capped with sheets of plastic and layers of clay to prevent toxic chemicals
and pollutants from seeping out. If these sites are left alone, they are often
colonized by weedy, exotic species. However, these eyesores can instead be
the focus of conservation efforts; planting native shrubs and trees attracts
birds and mammals that will bring in and disperse the seeds of a wide
range of native species. The Fresh Kills restored landfill site on Staten
Island in New York City is a good example of such restoration practices
(www.nycgovparks.org/park-features/freshkills-park). It occupies almost
1000 ha and has garbage mounds as tall as the Statue of Liberty, with a
volume 25 times that of the Great Pyramid of Giza. The landfill was closed
in 2001 and is now undergoing restoration to create a huge public park with
many elements of a native ecosystem, a project that is being implemented
in six phases to be completed by the year 2036. The eventual goal is to
create a large public parkland area (almost three times the size of New
York City’s Central Park) with abundant wildlife and many recreational,
cultural, and educational amenities.

Restoration Using Organisms


Restoration is often limited in geographic scope due to the associated cost,
but there is a growing movement to consider ways to repair ecosystems
at grand scales, made possible in some cases by reestablishing
certain ecosystem dynamics. Introducing animals (and other
types of organisms such as bacteria or fungi) can accomplish
Animals and other what would not be logistically or financially possible otherwise.
organisms can facilitate Rewilding is a term first introduced by Michael Soulé in the mid-
restoration at scales that 1990s to describe the reintroduction of top carnivores in order
might otherwise not be to regulate the system from the top down. Since that time, the
possible. term rewilding has been used in a variety of contexts, especially
attempts to restore aspects of ecosystems that last existed in the
Pleistocene era, more than 11,000 years ago. The most famous
of these is the reintroduction of wolves in Yellowstone National
Park (see Chapter 7). The idea that the restoration of an entire ecosystem
can be facilitated by the reintroduction of one or more missing functional
groups of animals is an approach that has also been referred to as trophic
rewilding (Svenning et al. 2015).
One place where trophic rewilding has been successful is the Oost-
vaardersplassen (“eastward-sailing wetland”), a nature reserve covering
about 56 km2 in a densely populated area just 52 km outside Amsterdam,
the capital of the Netherlands (www.staatsbosbeheer.nl/English). The Eu-
ropean landscape has arguably been more altered by human activities than
any other in the world; no forest, river, grassland, wetland, or almost any
other ecosystem of Europe has escaped human influence. But Frans Vera, a
Restoration Ecology 347

Figure 10.6 Reintroduced horses and other large herbivores in the Oost-
vaardersplassen helped decrease the dominance of trees through grazing.
(Photograph by Richard Primack.)

Dutch government scientist, believed that the reintroduction of large herbi-


vores that had been absent from the landscape for hundreds of years could
return the Oostvaardersplassen to a former, more functional state. Because
many of these species are now extinct, he introduced modern mammals as
ecological surrogates; beginning in the 1980s, he brought in Heck cattle in
place of extinct aurochs (wild cattle) and Konik ponies in place of tarpans,
the last of Europe’s wild horses (Figure 10.6). Vera also reintroduced red
deer, which were among the original herbivores in the area.
The rewilding effort had remarkable effects on the landscape. Popu-
lations of horses and deer exploded, grasslands and marshes began to
thrive as woody vegetation retreated, and many endangered birds took
up residence in the newly opened habitat. In 2013, the carcass of a wolf
was found in the Netherlands, marking their appearance for the first time
since the nineteenth century. However, the lack of significant predation
has led to booming herbivore populations that have no opportunity to
expand beyond the isolated reserve to other areas in search of food. Photos
and videos of starving animals were shown on television and in other
media, and people objected to such cruel “treatment” of animals, even
though it was a natural process (Economist 2013), leading to new manage-
ment policies of shooting suffering animals. There has also been public
debate over why managers are waiting for top predators to arrive on their
own rather than reintroducing them. Future plans for the reserve include
developing corridors to connect it to other nature reserves, as a part of a
348 Chapter 10

European network of protected areas, to facilitate the natural expansion


of predators and allow for migration of the herbivores. Although public
opinion has sometimes been critical, the proximity of the reserve to a met-
ropolitan area with 2.5 million people has enormous educational value,
allowing visitors to witness an ecosystem beginning to function as it did
thousands of years ago. More modest urban and suburban restoration
projects can serve similar purposes, but by European standards, this is a
major accomplishment.
Another type of restoration by animals is the release of biological con-
trol and bioremediation organisms. Unlike rewilding, these biological in-
troductions do not help restore ecological balance by mimicking historical
conditions, but rather are used to remove unwanted elements that were
introduced by humans. Bioremediation is the use of an organism to clean
up pollutants, such as prokaryotes that break down the oil in an oil spill
or wetland plants that take up agricultural runoff to clean the water (see
the section “Ecosystem services” in Chapter 3), whereas biological control
(also known as biocontrol) is the use of one type of organism, such as an
insect, to manage another, undesirable, species, such as an invasive plant.
Historically, a focus on human needs has led to problems in some cases,
when the released organism itself became a pest (see Chapter 4). However,
these experiences have informed a broader view that considers the whole
ecosystem, facilitating the use of both bioremediation and biocontrol in
restoration contexts with conservation-oriented goals (Seastedt 2014).
One example of the use of biocontrol for ecological restoration is the
release of the tamarisk leaf beetle (Diorhabda spp.) along rivers in the
western United States. Rivers and their riparian plant communities are
frequently the object of restoration efforts, but the geographic scale of the
problems often limits what can be done (Gonzalez et al. 2015). The focus of
restoration efforts in Texas, New Mexico, Arizona, and other western states
has frequently been the removal of exotic tamarisk (Tamarix spp.) trees,
which now dominate many riparian zones. When it behaves invasively,
this species is associated with a host of problems that negatively affect
both plants and animals (Sher 2013). Efforts at removing the tree with
bulldozers or herbicides risk harm to native species, are difficult to use
in remote regions, and are too expensive to implement on a large scale. In
response to these problems, more than a decade of research on biological
control of the tamarisk eventually led government scientists to release
the beetle in the wild in 2003 (Bean et al. 2013). By 2015, the beetle had
spread to cover hundreds of miles of rivers, feeding on tamarisk leaves
and turning acres of the invasive tree brown. The goal was to facilitate
the recovery of native trees and other species, and there is evidence that
this is occurring in some locations (Figure 10.7). However, it may be too
early to determine the long-term response of the ecosystem, especially in
the context of climate change (Hultine et al. 2015).
Just as the rewilding projects are not without problems, in this case
there have been criticisms about unintended effects of the biological
control on wildlife, particularly herpetofauna (reptiles and amphibians)
Restoration Ecology 349

(A) 2006 (B) 2013

(C) 2015

Figure 10.7 (A) Tamarisk (salt cedar, Tamarix) was


introduced from Eurasia in the 1800s and has since
formed monoculture thickets along rivers, as shown here
in a 2007 photograph of the Colorado River, from Fossil
Point near Moab, Utah. The restoration of such inacces-
sible places is being accomplished with the assistance of a
biological control insect, the tamarisk leaf-beetle (Dio-
rhabda spp.), which feeds on tamarisk leaves. (B) Stretches
of tamarisk trees turned brown after attack by the beetle, as
shown in a photograph taken in 2013. (C) This defoliation of the leaves allows the
restoration of native species, such as the green willows shown in this photograph
taken in 2015, which are colonizing along the riverbank and growing up through
the dead branches of the tamarisk tree. (A, Photograph by Anna Sher; B, Photo-
graph by Wayne Ranney; inset, courtesy of Eric Coombs, Oregon Department of
Agriculture, Bugwood.org; C, Photograph by Wright Robinson.)

(Bateman et al. 2014) and some species of birds (Sogge et al. 2013). Con-
cern for a federally listed bird that nests in the tamarisk even resulted in
litigation by an environmental group against the agency that released the
beetle. Several scientists have argued that the overall ecological benefits
of reducing the tamarisk are worth such problems (Tamarisk Coalition
350 Chapter 10

2016). This case illustrates the point that the benefits of any
In some cases, restoration restoration effort must always be weighed against perceived,
may be inadvisable due potential, and actual costs (Hinz et al. 2014).
to economic costs or A frequent scientific critique of both rewilding and biocon-
possible negative impacts trol restoration projects is that the practitioners spend too much
on the ecosystem. of their resources in active conservation and not enough on
monitoring, researching, or publishing findings. This problem
is caused in large part by the limited funding these projects
receive; they are often chronically underfunded and rely on
volunteers and nonprofit support. The lack of scientific publications gen-
erated by the world’s handful of rewilding projects in particular may be
a reason why they have not gained wider publicity and acceptance. Even
though the pace is slow, long-term efforts like these will provide important
lessons for restoration efforts elsewhere.

Moving Targets of Restoration


Since ecosystems change over time in response to climate change, plant
succession, the varying abundance of common species, and other factors,
the goals of restoration may have to be changed over time as well or modi-
fied to include temporal dynamics to remain realistic. In many situations
in which human activities have drastically altered the environment, some
biologists say we will have to accept novel ecosystems, in which there is
a mixture of native and nonnative species coexisting in a community un-
like the original or reference site (Hobbs et al. 2013; Kueffer and Kaiser-
Bunbury 2014). These novel ecosystems may differ in species composition
and function compared to historical conditions, reflecting the shifting na-
ture of species, the alteration of the environment, and even human values
(Harris et al. 2006).
Restoration ecology is increasingly addressing the issue of the moving
target, especially as it becomes evident that so many ecosystems simply
cannot be restored in the traditional sense (Arthington et al. 2014). The soil
chemistry or water availability may be too different or the elimination of an
introduced species may be impractical or even undesirable if that species
can perform an ecological role similar to that of a missing native species.
For example, research on native versus novel forests in Hawaii found that
species richness was greater in the novel forest that included both native
and exotic species and that total plant biomass and nutrient cycling were
either the same or greater there than in the native forest (Mascaro et al.
2012). Even though the native plant species in this system are declining,
proponents of novel ecosystems may not consider the novel forests for res-
toration efforts. In response to such perspectives, other researchers argue
that the concept of novel ecosystems may be mistakenly applied when the
barriers are social or political rather than ecological, and that projects may
use novelty as an excuse to not attempt complete restoration when it would
have otherwise been possible (Murcia et al. 2014; Simberloff et al. 2015).
Restoration Ecology 351

Restoration of Some Major Communities


In addition to rivers, many efforts to restore ecosystems have focused on
wetlands, lakes, prairies, and forests. These environments have been se-
verely altered by human activities and are good candidates for restoration
work.

Wetlands
Some of the most extensive restoration work has been done on wetlands,
including swamps and marshes (Halpern et al. 2007) (Figure 10.8). Be-
cause of wetland protection under the Clean Water Act and the US gov-
ernment policy of no net loss of wetlands, large development projects that
damage wetlands must repair them or create new wetlands to compensate
for those damaged beyond repair (Robertson 2006). The focus of these ef-
forts has been on re-creating the natural hydrology of the area and then
planting native species (Brinson and Eckles 2011). Many successful resto-
ration projects have resulted from this legislation; however, it has fallen
short of expectations due to a lack of monitoring and oversight (Clare
and Creed 2014). Strategies to restore the biodiversity of rivers include the
complete removal of dams and other structures and controlled releases
of water from dams (Helfield et al. 2007). For peatlands degraded by
harvesting for horticultural peat, a well-recognized restoration approach
is the moss layer transfer technique, which consists of spreading native
plant material collected from the top 10 cm of natural peatlands over the
restored area to facilitate reestablishment (Rochefort and Lode 2006).
Wetland restoration is motivated by more than just a concern for bio-
diversity, however. The 2005 destruction of New Orleans and other Gulf
Coast cities by Hurricane Katrina, and to a lesser extent by Hurricane
Rita soon after Katrina, was in part a result of the loss due to the devel-
opment of the region’s wetlands, which had protected the coast from the
force of hurricanes. The ensuing natural disaster has become a classic
example of the importance of ecosystem services to biological and human
communities alike (see Chapter 3). Ironically, the damage that followed
these hurricanes had been predicted seven years earlier by the Louisiana
Coastal Wetlands Conservation and Restoration Task Force (1998), which
had stressed the urgent need for immediate action to restore lost wetlands.
Restoration projects have begun, but if they are not adequately funded and
large enough in scope, New Orleans will remain vulnerable to another
destructive flood.
Experience has shown that efforts to restore wetlands often fail to close-
ly match the species composition or hydrologic characteristics of reference
sites. The subtleties of species composition, water movement, and soils, as
well as the site history, can be too difficult to match. Often the restored
wetlands are dominated by exotic, invasive species. However, the restored
wetlands usually do have some of the wetland plant species, or at least
similar ones, and can provide some of the functions of the reference sites
(Meyer et al. 2010). The restored wetlands also have some of the beneficial
352 Chapter 10

(A) 1973 Before drainage (B) 2000 After drainage

(C) 2005 After partial restoration

A
Permanent lake/reflooded marsh
Seasonal lake
Agriculture
Marsh vegetation

B
E
F
Iraq

50 km

Figure 10.8 Marsh restoration in Iraq as shown by Landsat images. The


images have false color, with marshland in red, agriculture in pink, and wetlands
in black and blue. (A) In 1973, marshes covered extensive areas of southern Iraq
and were home to about 400,000 Marsh Arabs. Three main marshes are labeled
1, 2, and 3. (B) As shown in the 2000 image, the marshes were drained by the
government for political reasons. (C) The reflooding of lakes and wetlands in re-
cent years has resulted in the restoration of some of the marsh vegetation, with
major restored areas indicated as A, B, and F. Other letters indicate sampling
sites. The new canals are still visible. (From Richardson and Hussain 2006; cour-
tesy of Curtis J. Richardson, Duke University Wetland Center.)
Restoration Ecology 353

ecosystem characteristics, such as flood control and pollution reduction,


and they are often valuable for wildlife habitat. Additional research into
restoration methods may result in further improvement.

Aquatic systems
Both freshwater and marine systems of all types are subject to degrada-
tion by pollution, overexploitation of resources, global warming, and other
factors (see Chapter 4), making them candidates for restoration. Aquatic
restoration may deal with problems regarding water chemistry, trophic
relationships with exotic species, and physical conditions of the shore or
bank. Although aquatic systems are often considered more resilient than
terrestrial systems, once damage has become severe, restoration can be
more complex.
One of the most common types of damage to lakes and ponds is cultural
eutrophication, or the accumulation of excess nutrients in the water caused
by human activity. Signs of eutrophication include an increased prevalence
of algal species (particularly surface scums of blue-green algae), decreased
water clarity and oxygen content, fish kills, and an eventual increase in
the growth of floating plants and other water weeds (see Figure 4.23). In
many lakes, the eutrophication process can be reversed by reducing the
amounts of mineral nutrients entering the water through better sewage
treatment or by diverting polluted water. One of the most dramatic and
expensive examples of lake restoration has been the effort to restore Lake
Erie (Sponberg 2009). Lake Erie was the most polluted of the
Great Lakes in the 1950s and 1960s, suffering from deteriorat-
ing water quality, extensive algal blooms, oxygen depletion in Lake restorations help
deeper waters, declining indigenous fish populations, and col- to improve water quality
lapsed commercial fisheries. To address this problem, the gov-
and restore the original
ernments of the United States and Canada have invested billions
species composition and
of dollars since 1972 in wastewater treatment facilities, reducing
community structure.
the annual discharge of phosphorus into the lake from 15,000
tons in the early 1970s to around 2000 tons today (International
Joint Commission 2014).
Many of the issues associated with lake restoration apply equally well
to marine ecosystems. These include shorelines, coral reefs, saltmarshes,
and mangroves. A number of large-scale projects are restoring estuar-
ies and bays damaged by human activities, including the Chesapeake
Bay in the eastern United States (Figure 10.9). Chesapeake Bay is one of
the most important fishing grounds and recreational areas in the United
States. However, pollution from residential, agricultural, and industrial
lands bordering the bay has caused a dramatic decline in the water qual-
ity, which affects all aspects of biodiversity. The economic consequences
of this pollution have also been apparent: harvests of fish and shellfish
have declined and the water has become unsafe for swimming. This type
of general pollution from an entire landscape is referred to as nonpoint
source pollution, and it requires a comprehensive restoration approach as
no single source of the pollution can be readily identified and contained.
354 Chapter 10

(A)

(B) Cumulative project costs (millions US$)


0 20 40 60 80 100 120

Stream and river


restoration 3000 stream
and river
Water treatment
projects
Bank stabilization
In-stream habitat
improvement $100 million
spent on
Channel reconfiguration
New York water treatment
Dam removal/retrofit projects
Fish passage
Chesapeake Bay
Other Pennsylvania
Watershed
Aesthetics/recreation/ New
management Jersey
Storm water management West D.C.
In-stream species Virginia Delaware
management Maryland
Flow modification
Virginia Chesapeake Number
Land acquisition
Bay of projects
Floodplain reconnection Costs

0 500 1000 1500 2000 2500 3000 3500


Number of projects
Figure 10.9 (A) A variety of measures have been taken to restore the health
of the Chesapeake Bay ecosystem. (B) This graph shows the cumulative costs
for each type of project and the number of projects. Stream and river restora-
tion is the most common type of project, and the most money has been spent
on water treatment projects. The map shows the watersheds that drain into the
bay. (A, © Mary F. Calvert/MCT/Getty Images; B, after Hassett et al. 2005.)
Restoration Ecology 355

In 1987 the federal, state, and local government bodies responsible for the
bay signed an agreement to reduce nutrient and sediment loads coming
into the bay by 40%, to be achieved mainly through improving the health
of streams and watersheds feeding water in. Since that time, over 4700
individual restoration projects have been implemented at a total cost of
over $400 million (Stokstad 2009). The largest number of projects involve
stream and river restoration, which includes regrading slopes and plant-
ing native vegetation. However, the most money has been spent on water
treatment projects.
A review of 235 studies of marine restoration projects found that al-
though millions of dollars are being spent on these projects, success has
been related more to the type of ecosystem (salt marshes and coral reefs
had greatest organism survival rates), site selection, and techniques rather
than to the amount of money spent (Bayraktarov et al. 2015). It also found
that most projects were short-lived and poorly monitored. For example, a
major weakness of the Chesapeake Bay restoration project was that only
5% of these projects have been monitored, and mainly just for vegetation
structure, and even fewer have been monitored for water quality to de-
termine whether they have been achieving the original goal of reducing
nutrient and sediment loads. This and other projects demonstrate that,
while society has accepted the need to restore large aquatic ecosystems,
scientists need to do a better job of ensuring that projects deliver the
services as promised.

Prairies and farmlands


Because they are species-rich, have many beautiful wildflowers, and can
be established within a few years, prairies represent ideal subjects for
restoration work (Foster et al. 2009). Many techniques have been used in
attempts at prairie restoration, but the basic method involves site prepara-
tion by shallow plowing, burning, and raking if prairie species are pres-
ent, or by eliminating all vegetation by plowing or applying herbicides
if only exotics are present. Native plant species are then established by
transplanting them in prairie sods obtained elsewhere, planting individu-
als grown from seed, or scattering prairie seeds collected from the wild or
from cultivated plants. The simplest method is to gather hay from a native
prairie and spread it on the prepared site. Native species are more likely
to become established in the absence of fertilizer, which tends to favor
nonnative species. (Of course, reestablishing the full range of plant spe-
cies, soil structure, and invertebrates could take centuries or might never
occur.) The Chicago metropolitan area is particularly well known for such
projects; some involve creating prairie grasslands with native prairie spe-
cies, rather than lawns, in suburban neighborhoods, while others involve
converting forests back to prairies (Figure 10.10).
A conundrum for restoration ecology, illustrated by work in the prairies,
is to determine the target state of the ecosystem; humans have had an
356 Chapter 10

Figure 10.10 (A) In the late (A)


1930s, members of the Civilian
Conservation Corps (one of the
organizations created by Presi-
dent Franklin Roosevelt in order
to boost employment during the
Great Depression) participated in
a University of Wisconsin project
to restore the wild species of a
midwestern prairie. (B) The prairie
as it looks today. (A, photograph
courtesy of the University of Wis-
consin Arboretum and Archives;
B, photograph courtesy of Molly
Field Murray.)

(B)

impact on many ecosystems for centuries or even millennia. For example,


early humans hunted many North American mammals to extinction more
than 12,000 years ago. Should North American grasslands be restored to a
state resembling those that existed either before European colonization, a
few hundred years ago, or those that existed before human colonization,
more than 12,000 years ago? One of the most ambitious proposed restora-
tions involves re-creating a short-grass prairie ecosystem, or “buffalo com-
mons,” on about 380,000 km2 of the Great Plains states, from the Dakotas
to Texas and from Wyoming to Nebraska (Adams 2006). Some of this land
is currently used for environmentally damaging and often unprofitable
agriculture and grazing supported by government subsidies. The human
Restoration Ecology 357

population of this region is declining as farmers and townspeople go out


of business and young people move away. From the ecological, sociologi-
cal, and economic perspectives, the best long-term use of much of this
region might be as a restored prairie ecosystem. The human population
of the region could stabilize around nondamaging core industries such as
tourism, wildlife management, and low-level grazing by cattle and bison,
leaving only the best lands in agriculture. As mentioned in “Restoration
Using Organisms” earlier in this chapter, some have argued for a process
of North American rewilding whereby large game animals from Africa
and Asia, such as elephants, cheetahs, camels, and even lions, would be
released in an attempt to re-create the types of ecological interactions that
occurred in North America before humans arrived on the continent (Hay-
ward 2009). Both these proposed projects are controversial because many of
the farmers and ranchers in the region want to continue their present way
of life without alteration, and they tend to be highly resentful of unwanted
advice and interference from scientists and the government. The projects
are also controversial because of the proposed release of nonnative mam-
mals in North American ecosystems.

Tropical dry forest in Costa Rica


An exciting restoration process has been ongoing since 1985 in northwest-
ern Costa Rica. The tropical dry forests of Central America have long suf-
fered from large-scale conversion to cattle ranches and farms. This de-
struction has gone largely unnoticed, as international scientific and public
attention has focused on more glamorous rain forests elsewhere in South
and Central America (e.g., see Chapters 6 and 11). The American ecologists
Daniel Janzen and Winnie Hallwachs have been working with Costa Rica’s
Ministry of the Environment, the resident staff, and the Guanacaste Dry
Forest Conservation Fund to restore the biology and cultural connectivity
of 130,000 ha of land and 43,000 ha of overfished marine habitat in Area
de Conservación Guanacaste (ACG) (Ehrlich and Pringle 2008; www.ac-
guanacaste.ac.cr) (Figure 10.11).
The restoration of these marginal ranchlands, low-quality farms, and
forest fragments includes eliminating brush fires started by people, ban-
ning logging and hunting, and occasionally planting both native and exotic
trees to shade out introduced African grasses. Light livestock grazing ini-
tially reduced grass volume and then was phased out as the forest returned
through natural animal- and wind-borne seed dispersal. In 29 years, this
process has converted tens of thousands of hectares of pastures and old
fields to a species-rich, dense young forest, with abundant and growing
populations of native animals. Nonetheless, the area will require an esti-
mated 200–500 years to regain the original forest structure.
A key element in the restoration plan is what has been termed bio-
cultural restoration, meaning that ACG staff members teach basic biol-
ogy and ecology on-site to 2500 students in grades 4 through 6 from 53
358 Chapter 10

(A) (B)

(C)
Figure 10.11 The Area de Conservación
Guanacaste (ACG) is an experiment in restoration
ecology—an attempt to restore the devastated
and fragmented tropical dry forest of Costa Rica.
(A) A barren grassland with scattered forest frag-
ments was heavily grazed by cattle and frequently
burned. (B) Native trees and other species became
established once again in this young forest after
17 years without cattle and fire. Note the person
in the lower left for scale. (C) Daniel Janzen, an
ecologist from the United States, is a driving force
behind the restoration project in Guanacaste. Here
he explains a land purchase deal to the board of di-
rectors of the Guanacaste Dry Forest Conservation
Fund. (Photographs courtesy of Brad Zlotnick.)

neighboring schools and also give presentations to citizen groups, all as


part of the ACG core mission. This effort, combined with the fact that
all 95 members of ACG’s staff and administration are resident Costa
Ricans, has resulted in residents viewing ACG as if it were a large ranch
producing “wildland resources” for the community rather than an ex-
clusionary “national park.” Janzen (quoted in Allen 1988) believes that, in
rural areas such as Guanacaste Province, providing an opportunity for
learning about and from nature can be one of the most valuable functions
of conserved wildlands:
The goal of biocultural restoration is to give back to people the under-
standing of the natural history around them that their grandparents
had. These people are now just as culturally deprived as if they could
no longer read, hear music, or see color.
Restoration Ecology 359

This restoration effort has accomplished so much and has attracted


widespread attention in part because a highly articulate individual—Jan-
zen—commits most of his time and resources to a cause in which he pas-
sionately believes (Laurance 2008a). Janzen’s enthusiasm and vision have
inspired many other people to join his cause, and he is a classic example
of how one individual can be a potent force for conservation.

The Future of Restoration Ecology


E.O. Wilson predicted in 1992, “Here is the means to end the great extinc-
tion spasm. The next century will, I believe be the era of restoration in
ecology.” He may be right. Restoration ecology is an evolving and rapidly
growing discipline, with its own scientific society (the Society for Ecologi-
cal Restoration) and increasing numbers of journals: Restoration Ecology,
Ecological Restoration, The Journal of Ecosystem Restoration, Ecological Man-
agement and Restoration, and others. Ecosystems are being restored using
methods developed by the discipline, books are being written about the
subject, and new courses are being taught at more universities. Scientists
are increasingly able to make use of the growing range of published studies
and suggest improvements in restoration techniques. At its best, restored
land can provide new opportunities for protecting biodiversity and gen-
erating an appreciation of nature.
Conservation biologists in this field must take care to ensure
that restoration efforts are not simply public relations endeavors
taken by environmentally damaging corporations that are only Ecological restoration
interested in continuing business as usual (Maron et al. 2012). is an important and
A 5 ha “demonstration” or “best practices” project in a highly growing tool for
visible location does not compensate for thousands or tens of
conservation, but the
thousands of hectares damaged elsewhere and should not be
protection of existing
accepted as adequate by conservation biologists. Attempts to
biodiversity remains the
mitigate or offset the destruction of an intact biological com-
first priority.
munity by the building of a similar species assemblage at a new
location is almost certainly not going to provide a home for the
same species or provide similar ecosystem functions (Moreno-
Mateos et al. 2015a).
The best long-term strategy remains protecting and managing biologi-
cal communities where they are found naturally; many researchers argue
that the requirements for the long-term survival of all species are most
likely to be found there, and the protection of intact systems is gener-
ally cheaper and easier than repairing systems that have been degraded.
In addition, we need to consider restoring ecosystems in anticipation of
the impacts of climate change (see Chapter 8). There are many technical,
scientific, logistical, and economic challenges for future restoration ecolo-
gists and managers to address in our pursuit to repair damage we have
done as a species.
360 Chapter 10

Summary
„„Ecological restoration is the practice of tat need to be monitored to determine
reestablishing populations, ecosystems, whether they are reestablishing the
and landscapes that include degraded, composition of historical species and the
damaged, or even destroyed habitat. functions of the ecosystem.
Restoration ecology provides methods „„Biological control and bioremediation
for reestablishing species, whole biologi- are tools whereby organisms such as
cal communities, and ecosystem func- insects or protists can be used to re-
tions in degraded habitat. move invasive species and pollutants
„„The establishment of new communities (respectively).
on degraded or abandoned sites pro- „„In some cases, restoration to a former
vides an opportunity to enhance biodi- state is impractical or impossible due to
versity and can improve the quality of the nature or extent of the degradation,
life for the people living in the area. Res- the presence of invasive species, or cli-
toration ecology can also provide insight mate change. In such cases, a novel eco-
into community ecology by testing our system that may have some of the same
ability to reassemble a biological com- functionality of the original one may be
munity from its native species. considered, but it should not be valued
„„Restoration projects begin by eliminat- over the native ecosystem.
ing or neutralizing factors that prevent „„Creating new habitat to replace lost
the system from recovering. Then some habitat elsewhere, which is known as
combination of site preparation, habi- compensatory mitigation or biological
tat management, and reintroduction of offsetting, has value but should be re-
original species gradually allows the garded as only part of an overall conser-
community to regain the species and vation strategy that includes the protec-
ecosystem characteristics of designated tion of species and ecosystems where
reference sites. Attempts to restore habi- they naturally occur.

For Discussion
1. Restoration ecologists are improving 3. What do you think are some of the easi-
their ability to restore biological com- est ecosystems to restore? The most dif-
munities. Does this mean that biological ficult? Why?
communities can be moved around the 4. Aldo Leopold encouraged humans to
landscape and positioned in convenient “keep every cog and wheel” in order to
places that do not inhibit the further ex- maintain healthy ecosystems. Is it neces-
pansion of human activities? sary, or even possible, to return every
2. What methods and techniques could you missing species back into a restored
use to monitor and evaluate the success ecosystem?
of a restoration project? What timescale
would you suggest using?
Restoration Ecology 361

Suggested Readings
Cole, I. A., S. M. Prober, I. D. Lunt, and T. B. Koen. 2016. A plant traits ap-
proach to managing legacy species during restoration transitions in tem-
perate eucalypt woodlands. Restoration Ecology. doi: 10.1111/rec.12334. It is
important to preserve desirable species while removing undesirable ones.
Corlett, R. T. 2016. Restoration, reintroduction and rewilding in a changing
world. Trends in Ecology and Evolution. doi: 10.1016/j.tree.2016.02.017.
Dodds, W. K., K. C. Wilson, R. L. Rehmeier, G. L. Knight, S. Wiggam, J. A.
Falke, and 2 others. 2008. Comparing ecosystem goods and services pro-
vided by restored and native lands. BioScience 58: 837–845. Within 10 years
of restoration, restored ecosystems provide 31%–93% of the benefits of
native lands.
Felson, A. J., M. A. Bradford, and T. M. Terway. 2013. Promoting Earth stew-
ardship through urban design experiments. Frontiers in Ecology and the
Environment 11: 362–367. Cities can be restored to reduce environmental
impacts and create a healthier environment for people.
Fraser, L. H., W. L. Harrower, H. W. Garris, S. Davidson, P. D. N. Hebert, R.
Howie, and 8 others. 2015. A call for applying trophic structure in ecologi-
cal restoration. Restoration Ecology 23: 503–507. The traditional focus of
restoration ecology on vegetation must be challenged.
Galatowitsch, S. M. Ecological Restoration. 2012. Sinauer Associates, Sunder-
land, MA. A comprehensive overview of the strategies being used around
the world to reverse human impacts to landscapes, ecosystems, and
species.
González E., A. A. Sher, E. Tabacchi, A. Masip, and M. Poulin. 2015. Restora-
tion of riparian vegetation: A global review of implementation and evalu-
ation approaches in the international, peer-reviewed literature. Journal of
Environmental Management 158: 85–94. The restoration of riverbank com-
munities can have a wide range of goals and utilize a variety of method-
ological approaches.
Handel, S. N. 2016. Greens and greening: Agriculture and restoration ecol-
ogy in the city. Ecological Restoration 34(1): 1–2. Urban restoration provides
many benefits to both wildlife and people.
Keenelyside, K., N. Dudley, S. Cairns, C. Hall, and S. Stolton. 2012. Ecological
restoration for protected areas:principles, guidelines and best practices (Vol. 18).
IUCN. A best practices guide based on input from restoration ecologists
from more than a dozen countries.
Kueffer, C. and C. N. K. Kaiser-Bunbury. 2014. Reconciling conflicting perspec-
tives for biodiversity conservation in the Anthropocene. Frontiers in Ecology
and the Environment 12: 131–137. Restored landscapes and novel ecosys-
tems with mixtures of native and nonnative species might be best suited to
survive increasing human impacts.
Oppenheimer, J. D., S. K. Beaugh, J. A. Knudson, P. Mueller, N. Grant-Hoff-
man, A. Clements, and M. Wight. 2015. A collaborative model for large-
scale riparian restoration in the western United States. Restoration Ecology
23(2): 143–148. Federal agencies and nonprofit organizations can work
together to facilitate restoration.
The Challenges
11 of Sustainable
Development
Sustainable Development at the International Approaches to
Local Level 365 Sustainable Development 374
Conservation at the National Funding for Conservation 383
Level 372

Growing plants on rooftops (green roofs) helps


pollinators, absorbs CO2, and can even provide
habitat in what is otherwise an urban desert.
This rooftop garden in Japan provides vegetables
and fruits for local residents.
E
fforts to preserve biological diversity some-
times conflict with both real and perceived
human needs (Figure 11.1). Increasingly,
many conservation biologists, policymakers, and land
managers are recognizing the need for sustainable
development—economic development that satis-
fies both present and future needs for resources and
employment while minimizing the impact on biodi-
versity and functioning ecosystems (Selomane et al.
2015). Sustainable development, a term sometimes
used interchangeably with sustainability, can be con-
trasted with more typical development that is unsus-
tainable. Unsustainable development cannot con-
tinue indefinitely because it destroys or uses up the
resources on which it depends. As defined by some
environmental economists, economic development
implies improvements in the efficiency, organization,
and distribution of resource use or other economic
activity but not necessarily increases in resource
consumption. Economic development is clearly distin-
guished from economic growth, which is defined as
material increases in the amount of resources used.
Sustainable development is a useful and important
concept in conservation biology because it empha-
sizes improving current economic development and
limiting unsustainable economic growth.
364 Chapter 11

Figure 11.1 Sustainable development seeks to address the conflict that


exists between development to meet human needs and the preservation of the
natural world. (Top photograph © FloridaStock/Shutterstock; bottom photograph
© kavram/Shutterstock.)

Investment in national park infrastructure that improves the


The goal of sustainable protection of biodiversity and provides revenue opportunities for
economic development is
local communities is an example of movement toward sustain-
able development, as is the implementation of less destructive
to provide for the current
logging and fishing practices. Especially in developing countries,
and future needs of
sustainability is tightly intertwined with issues of social justice;
human society while at
one cannot have sustainable development without ensuring that
the same time protecting
all parties benefit. This is true in part because environmental
species, ecosystems, degradation disproportionately affects the poor. It has been
and other aspects of argued that the combination of environmental, economic, and
biodiversity. social considerations explicit in sustainable development will
inevitably lead to improved quality of life, corporate profits, and
human health in addition to environmental benefits (Dernbach
and Cheever 2015).
Unfortunately, the term sustainable development has become overused
and Introduction
is often misappropriated.
to Conservation BiologyFew politicians or businesses are willing to
1E Primack/Sher
proclaim
Sinauerthemselves
Associates to be against sustainable development. Thus, many
Morales Studio/SA
Primack_Sher1E_11.01 Date 3-2-16 3-9-16
The Challenges of Sustainable Development 365

large corporations, and the policy organizations that they fund, misuse
the notion of sustainable development to “greenwash” their industrial ac-
tivities, with only limited change in actual practice. In these cases, both
ecosystems and people will likely suffer.
For instance, a plan to establish a huge mining complex in the middle
of a forest wilderness cannot justifiably be called sustainable development
simply because a small percentage of the land area is set aside as a park.
Not only will precious habitat be lost, but local inhabitants who depend
on resources from that forest will be impacted as well. Waste from the
mine can poison fish and people alike. Similarly, building huge houses
filled with “energy-efficient” appliances and cars that boast the latest
energy-saving technology but are routinely driven long distances cannot
really be called sustainable development or “green technology” when the
net result is increased energy use. Alternatively, some people champion
the opposite extreme, claiming that sustainable development means that
vast areas must be kept off-limits to all development and should remain
as, or be allowed to return to, wilderness. This may not be the best op-
tion for either conservation goals or people: some places require active
restoration (see Chapter 10) or active management (see Chapters 8 and 9)
to best protect biodiversity, and barring all people will inevitably harm
local interests.
The primary conflict of sustainable development is often not between
people and nature so much as it is between the powerful and the vulner-
able. As with all such disputes, informed scientists and citizens must study
the issues carefully, identify which groups are advocating which positions
and why, and then make careful decisions that best meet both the needs
of human society and the protection of biodiversity and ecosystems. In
many cases this involves compromise, and in most cases compromises
form the basis of government policy and laws, with conflicts resolved by
government agencies and the courts.

Sustainable Development at the Local Level


Most efforts to find approaches that promote both the preservation of spe-
cies and habitats and the needs of society rely on initiatives from concerned
citizens, conservation organizations, and government officials. These ef-
forts may take many forms, but they begin with individual and group
decisions to prevent the destruction of habitats and species in order to
preserve things of perceived economic, cultural, biological, scientific, or
recreational value. The results of these initiatives often end up codified
into environmental regulations or laws.

Local and regional conservation regulations


In modern societies, local (city and town) and regional (county, state, and
provincial) governments pass laws to provide protection for species and
366 Chapter 11

habitats while at the same time allowing development for the continued
needs of society. Often, but not always, these local and regional laws are
comparable to, or stricter than, national laws, particularly for protections
of clean water and air, and less often for endangered species. Such laws
are passed because citizens and political leaders feel that they represent
the will of the majority and provide long-term benefits to society. The most
prominent of these laws govern when and where hunting and fishing can
occur; the size, number, and species of animals that can be taken; and the
types of weapons, traps, and other equipment that can be used. Restrictions
are enforced through licensing requirements and patrols by game wardens
and police. In some settled and protected areas, hunting and fishing are
banned entirely. Similar laws affect the harvesting of plants, seaweed, and
shellfish. Certification of the origin of biological products may be required
to ensure that wild populations are not depleted by illegal collection or har-
vest. These restrictions have long applied to certain animals such as trout
and deer and to plants of horticultural interest such as orchids, azaleas,
and cacti. More recently, there are certification programs for the origin of
ornamental fish, timber, and other products.
Laws that control the ways in which land is used are another means
of protecting biodiversity (Reed et al. 2014). For example, on a more local
scale, vehicles and even people on foot may be restricted from habitats and
resources that are sensitive to damage, such as birds’ nesting areas, bogs,
sand dunes, wildflower patches, and sources of drinking water. Uncon-
trolled fires may severely damage habitats, so practices, such as campfires,
that contribute to accidental fires are often rigidly controlled. Zoning laws,
among the strongest and most widely used restrictions, sometimes prevent
construction in sensitive areas such as barrier beaches and floodplains.
Even where development is permitted, building permits are often reviewed
carefully to ensure that damage is not done to endangered species or eco-
systems, particularly wetlands. For major regional and national projects,
such as dams, canals, mining and smelting operations, oil extraction, and
highway construction, environmental impact statements must be prepared
that describe the damage that such projects will or could possibly cause so
that these projects can be conducted in a more environmentally sensitive
manner.
One of the most powerful strategies in protecting biodiversity at the
local and regional levels is the designation of intact biological communi-
ties as nature reserves, conservation land, and state and provincial parks
and forests (see Chapter 8). Government bodies buy land and establish
protected areas for various uses—local parks for recreation, conservation
areas to maintain biodiversity, forests for timber production and other
uses, and watersheds to protect water supplies.
The passage and enforcement of conservation-related laws on a local
level can become an emotional experience that divides a community and
can even lead to violence. To avoid such counterproductive outcomes, con-
servationists must be able to convince the public that using resources in a
The Challenges of Sustainable Development 367

thoughtful and sustainable manner creates the greatest long-term benefit


for the community. The general public must be made to look beyond the
immediate benefits that can come with the rapid and destructive exploita-
tion of resources. For example, towns often need to restrict development
in watershed areas to protect water supplies; this may mean that houses
and businesses are not built in these sensitive areas and landowners may
have to be compensated for these lost opportunities. It is essential that
conservation biologists clearly communicate the reasons for restrictions
that protect biodiversity and ecosystems. Those affected by the restrictions
can become allies in the protection of resources if they understand the
importance and long-term benefits of reduced access. These people must
be kept informed and consulted throughout the decision-making process.
Conservation biologists must develop the necessary skills, combined with
the best science, to negotiate and compromise; to encourage conservation
actions in others and understand their perspectives; and to explain posi-
tions, regulations, and restrictions. These skills are part of the growing
field of conservation psychology (Clayton and Myers 2015; Van Vugt 2009).
Having a fervent belief in one’s cause is no longer enough.

Land trusts and related strategies


In many countries, nonprofit, private conservation organizations are among
the leaders in acquiring land for conservation (Bode et al. 2011). In the Neth-
erlands, about half the protected areas are privately owned. In
the United States, over 15 million ha of land are protected at the
local level by about 1700 land trusts, which are private, nonprofit
corporations established to protect land and natural resources Land trusts are
(www.landtrustalliance.org). At a national level, major orga- private conservation
nizations such as The Nature Conservancy and the National organizations that
Audubon Society have protected an additional 10 million ha purchase and protect
in the United States (Figure 11.2). While the purchase of land land. Conservation
may seem the most straightforward approach to conservation, easements and limited
in practice, property law can be quite complex and may differ development agreements
considerably between countries. are also used by land
Land trusts are common in Europe. In Britain, the National trusts to increase the
Trust has more than 3.8 million members and 62,000 volunteers amount of land under
and owns about 250,000 ha of land, much of it farmland, includ- protection.
ing 57 National Nature Reserves, 466 Sites of Special Scientific
Interest, 355 Areas of Outstanding Natural Beauty, and 40,000
archaeological sites. The Royal Society for the Protection of Birds
has more than 1 million members and manages 200 reserves with an area
of almost 130,000 ha. A major emphasis of many of these reserves is nature
conservation and education, often linked to school programs. These private
reserve networks are collectively referred to as Conservation, Amenity,
and Recreation Trusts (CARTs), a name that reflects their varied objectives.
In addition to purchasing land outright, both governments and conser-
vation organizations protect land through conservation easements (CEs),
368 Chapter 11

Figure 11.2 Land trusts may own and manage land or may give it to local
or national governments in special agreements. Here, the chief operating of-
ficer of The Nature Conservancy signs over 10 acres to establish the Everglades
Headwaters National Wildlife Refuge. (Photograph by Tom MacKenzie, USFWS,
Jan. 18, 2012, at the FFA training facility in Haines City, FL, about 50 miles
south of Orlando.)

also called conservation covenants, in which landowners give up the right


to develop, build on, or subdivide their property, typically in exchange for
a sum of money, lower real estate taxes, or some other tax benefit (Farmer
et al. 2011; see Chapter 9). CEs can have a variety of goals that may or
may not explicitly include protection of species and/or habitats, which
have important implications for how they are managed and the result-
ing effectiveness for conservation (Figure 11.3). For many landowners,
accepting a conservation easement is an attractive option: they receive a
financial gain while still owning their land and are able to feel that they
are assisting conservation objectives. In general, landowners are most
willing to consider CEs when they are well paid; the agreement is short
term, lasting just a few years rather than permanent; and there is no legal
contract involved (Sorice et al. 2013). Of course, the offer of lower taxes or
money is not always necessary; many landowners will voluntarily accept
conservation restrictions without compensation.
Another strategy that land trusts and local and regional governments
use is limited development, also known as conservation development
(Milder et al. 2008). In these situations, a landowner, a property devel-
oper, and a government agency and/or conservation organization reach
The Challenges of Sustainable Development 369

100

80
Percentage

60
Prohibited
Does not specify
40
Permitted

20

0
t t
en es ng in
g
tio
n
io
n
in
g
rv zi is in
pm a ra ar
m ea v
el
o rh G F ecr b di M
ev be a lr Su
D Ti
m ci
er
m
m
Co
Figure 11.3 Conservation easements are intended to protect land that
might otherwise be developed; however, a large proportion of them do allow
some forms of development and other uses. These figures reflect a review of 269
CE documents that spanned six US states. When activities were not specified,
they were likely to be allowed by default. (After Rissman et al. 2015.)

a compromise that allows part of the land to be commercially developed


while the remainder is protected by a conservation easement. Limited
development allows the construction of necessary buildings and other
infrastructure for an expanding human society; the projects are often
successful precisely because being located adjacent to conservation land
enhances the value of the developed land.
Governments and conservation organizations can further encourage
conservation on private lands through other mechanisms, including com-
pensating private landowners for desisting from some activity that dam-
ages the environment and for implementing conservation activity (Knoot
et al. 2015). Conservation leasing involves providing payments to private
landowners who actively manage their land for biodiversity protection. Tax
deductions and payments can also be obtained for any costs of restoration
or management, including weeding, controlled burning, establishing nest
holes, and planting native species. In some cases, private landowners may
still be allowed to develop their land later, even if endangered species
come to live there.
A related idea is conservation banking, in which a landowner delib-
erately preserves an endangered species or a protected habitat type such
as wetlands, or even restores degraded habitat and creates new habitat.
A conservation bank is like a financial bank in that it is intended to
be a stable protector, but instead of money, it protects natural resource
values. The US Fish and Wildlife Service (USFWS) evaluates proposals
370 Chapter 11

for new conservation banks “in the context of unavoidable impacts of


proposed projects to listed species” (USFWS 2003). Once approved, the
USFWS awards a landowner habitat or species credits in exchange for
permanently protecting the land and managing it for these species. These
credits can then be purchased by developers in compensation or as a
biodiversity offset for a similar habitat that is being destroyed elsewhere
by a construction project (Pilgrim et al. 2013). For example, the Muddy
Boggy Conservation Bank in eastern Oklahoma was expanded to nearly
230 ha in exchange for 522 credits for the endangered American burying
beetle (Nicrophorus americanus) (BusinessWire 2015). This conservation
bank was specifically established to assist Oklahoma industries such as
energy development, pipeline construction, and transportation projects
that might have otherwise been shut down if the endangered species was
found on the land. The industry can be relieved of the responsibility for
how their actions may affect the beetle by purchasing credits that are
then used to support the Muddy Boggy Conservation Bank’s ability to
protect the beetle there.
A related approach is the payments for ecosystem services (PES) pro-
gram (see Chapter 9), in which a landowner is paid for providing specific
conservation services (Figure 11.4) (Naeem et al. 2015). Utility companies
may also gain carbon credits by paying for habitat protection (e.g., paying
a landowner for not cutting down a forest) and restoration (e.g., paying a
landowner for planting trees and establishing a new forest); these carbon
credits are then used to offset the carbon emissions produced through the
burning of fossil fuels. On a larger scale, carbon offset payments by govern-

100 Controls
Protected
80
Nest survival over
nesting period (%)

60

40

20

0
Sarus crane Lesser adjutant

Figure 11.4 Fledgling success for nests of the Sarus crane (Grus antigone) and
lesser adjutant (Leptoptilos javanicus) that are protected by villagers in the north-
ern plains of Cambodia versus those that were not. Villagers participating in the
program were paid by the Wildlife Conservation Society, an international conser-
vation organization based in the United States. The local inhabitants were able to
significantly supplement their incomes with payments that provided extra incen-
tives for achieving successful nests. (After Clements et al. 2013. Crane photograph
© ScratchArt/Shutterstock; adjutant photograph © Arco Images GmbH/Alamy.)
The Challenges of Sustainable Development 371

ments and international corporations can be used to compensate


for emissions of greenhouse gases (Kiesecker et al. 2009). Conservation lands
Conservation concessions are an innovative approach in require ongoing manage-
which conservation organizations outbid extractive industries ment and continuous
such as logging companies, not for ownership of the land, but vigilance, but they
for the rights to use and protect it. The government or large often provide extensive
landowner receives the same annual income from a conser- benefits to the local
vation organization that would have been paid by a logging society and economy.
company, and the animals and plants of the area are protected
rather than destroyed.

Enforcement and public benefits


The conservation measures described in this section and elsewhere in
this book must be continuously monitored to make sure that regulations
and laws are enforced and that agreements are being carried out, particu-
larly in cases where destruction cannot be easily reversed (Rissman and
Butsic 2011). In one scenario, a developer may agree to limit the amount
of development and conserve an area of forest but then obtain construc-
tion permits, ignore the agreement, and clear all the trees. By the time
action can be taken to stop the developer, the trees and the habitat they
provided are gone and cannot be easily replaced. Even if sanctions such
as fines or forfeiting of bonds are imposed, the developer may feel that
the potential profits outweigh such considerations, and managers and
officials usually take a “what’s done is done” approach and allow the
cleared land to be developed. Conservation workers need to raise aware-
ness so that the public and the judicial system view “breach of promise”
against the environment with the same seriousness as similar crimes
against personal property.
The most powerful tool for enforcement has been fining violators, es-
pecially when these funds are directly applied to conservation, such as the
restoration of damaged land. When the terms of a conservation easement
are broken or other environmental harm is done, there may be a restoration
remedy; that is, the violator may be required to pay to return the degraded
property to as much of its original state as possible. For example, if a land
trust is obligated to preserve a stand of trees under the terms of a conser-
vation easement but then allows those trees to be cut down, the land trust
may be obligated to replant trees and nurture them to maturity at a cost
that well exceeds that of the land itself.
Public perception can also be a source of problems. Local efforts by land
trusts to protect land are sometimes criticized as being elitist because they
provide tax breaks only to individuals wealthy enough to take advantage of
them while decreasing the revenue collected from land and property taxes.
Other analysts argue that land used in other ways, such as for agriculture
or commercial activity, is more productive. Although land in trust may
initially yield lower tax revenues, the loss is often offset by the increased
value and consequent increased property taxes of houses and land adja-
372 Chapter 11

cent to the conservation area. In addition, the employment, recreational


activities, tourist spending, and research projects associated with nature
reserves and other protected areas generate revenue throughout the local
economy, which benefits local residents (Di Minin et al. 2013). Finally, by
preserving important features of the landscape and natural communities,
local nature reserves also protect and enhance the cultural heritage of the
local society, a consideration that must be valued if sustainable develop-
ment is to be achieved.

Conservation at the National Level


Throughout the modern world, national governments play a leading role
in conservation activities. The level of a government’s conservation actions
can substantially affect the conservation outcomes within its borders. Un-
intentionally triggering conflict with local government during conserva-
tion implementation and lack of government funding are two significant
problems faced by parties actively working to reestablish endangered spe-
cies (Crees et al. 2016). Conservation biologists contribute
to these efforts by providing government officials with
key information on threats to biodiversity, with the hope
and expectation that resulting laws and regulations will
be used to protect biodiversity (Figure 11.5).
Similar to local and regional governments, national
governments can use their revenues and authority to buy
land for conservation. In the United States, special funding
mechanisms exist at the national level, such as the Lands
Legacy Initiative and the Land and Water Conservation
Fund, to purchase land for conservation purposes.
The establishment of national parks is a particularly
important conservation strategy. National parks are the
single largest source of protected lands in many countries.
For example, Costa Rica’s national parks protect about
62,000 ha, or about 12% of the nation’s land area (www.
costarica-nationalparks.com). Outside the protected areas,
deforestation is occurring rapidly, and soon national
parks may represent the only undisturbed habitat and
source of natural products, such as timber, in the whole
country. The US National Park Service protects about 34
million ha with 410 sites. The United States government
Figure 11.5 A field biologist takes
a US congressman into a woodland to
also protects biodiversity in more than 598 National Wild-
explain how habitat fragmentation and life Refuges covering 61 million ha, and the U.S. Bureau
the loss of biodiversity contribute to of Land Management’s National Landscape Conservation
increasing rodent populations and the System has 873 sites covering 13 million ha, including
rising incidence of Lyme disease. (Pho- National Conservation Areas, National Monuments, and
tograph by J. Halpern.) Wilderness Areas.
The Challenges of Sustainable Development 373

National legislatures and governing agencies are the principal bodies


for developing policies that regulate environmental pollution. Laws are
passed by legislatures and then implemented in the form of regulations
imposed by government agencies. Laws and regulations affecting air emis-
sions, sewage treatment, waste dumping, and the development of wetlands
are often enacted to protect human health and property and resources
such as drinking water, forests, and commercial and sport fisheries. The
level of enforcement of these laws demonstrates a nation’s determination
to protect the health of its citizens and the integrity of its natural resources.
At the same time, these laws protect biological communities that would
otherwise be destroyed by pollution and other human activities. The air
pollution that exacerbates human respiratory disease, for instance, also
damages commercial forests and biological communities, and pollution
that ruins drinking water also kills terrestrial and aquatic species such as
turtles, fish, and aquatic plants.
National governments can also have a substantial effect
on the protection of biodiversity through the control of their National governments
borders, ports, industry, and commerce. To protect forests and protect designated
regulate their use, governments can ban logging, as was done endangered species
in Thailand following disastrous flooding; they can restrict the within their borders,
export of logs, as was done in Indonesia; and they can penal- establish national parks,
ize timber companies that damage the environment. Certain
and enforce legislation on
kinds of environmentally destructive mining can be banned.
environmental protection.
Methods of shipping oil and toxic chemicals can be regulated.
Conservation biologists can provide government officials with
key information for developing the needed policy framework,
and resource managers and others can then use the resulting
laws and regulations to protect biodiversity.
Despite the fact that many countries have enacted legislation to protect
endangered species, forests, wetlands, and other aspects of biodiversity, it
is also true that national governments are sometimes unresponsive to re-
quests from conservation groups to protect the environment. Governments
have even acted to remove the legal protected status of national parks,
sacred forests, and other conservation areas (degazettement; see Chapter
8), in order to facilitate the extraction of natural resources and economic
development (Hardy et al. 2016). Governments sometimes do this because
they feel that the needs of the broader regional and national society for
natural resources and economic development are more important than
the needs of the local people, the environment, or the ecosystem services
they may forgo. There are also cases where the downgrading, downsizing,
and degazettement of protected areas are linked to certain government
officials who profit personally from their actions. In some cases, national
governments recognize that local people are best able to protect ecosystems
close to where they live and have relinquished control of these resources
to local governments, village councils, and conservation organizations.
374 Chapter 11

International Approaches to
Sustainable Development
The biological diversity needed for humanity’s future well-being
International cooperation is concentrated in the tropical countries of the developing world,
and agreements to most of which are relatively poor and experiencing rapid rates
protect biodiversity are of population growth, development, and habitat destruction (see
needed for migratory Figure 6.20). Developing countries may be willing to preserve
species and for occasions biodiversity, but they are often unable to pay for the habitat pres-
when threats occur ervation, research, and management required for the task. The
across countries. developed countries of the world (including the United States,
Canada, Japan, Australia, and many European nations) must work
together with tropical countries to preserve the biodiversity need-
ed by the world as a whole.
While the major legal and policing mechanisms that presently exist in
the world are based within individual countries, international cooperation
to protect biodiversity is an absolute requirement for several reasons:
• Species migrate across international borders. Conservation efforts must
protect species at all points in their ranges; efforts in one country will
be ineffective if critical habitats are destroyed in a second country to
which an animal migrates (Ripple et al. 2014). For example, efforts to
protect migratory bird species in northern Europe will not work if the
birds’ overwintering habitat in Africa is destroyed.
• International trade in biological products is commonplace. A strong demand
for a product in one country can result in the overexploitation of the
species in another country to supply this demand.
• Biodiversity provides internationally important benefits. The community
of nations benefits from the species and genetic variation used in ag-
riculture, medicine, and industry; the ecosystems that help regulate
climate; and the national parks and other protected areas of interna-
tional scientific and tourist value. These benefits have been estimated
to be in the trillions of dollars (McCarthy et al. 2012).
• Many environmental pollution problems that threaten ecosystems are inter-
national in scope. Such threats include atmospheric pollution and acid
rain; the pollution of lakes, rivers, and oceans; greenhouse gas produc-
tion and global climate change; and ozone depletion (Lin et al. 2014).

International Earth summits


Given these realities, there is motivation to make progress on conserva-
tion issues by bringing together leaders at international meetings. There
have been six of these meetings that consider environmental issues in the
context of sustainable development:
• UN Conference on the Human Environment (1972)
• World Commission on Environment and Development (1987)
The Challenges of Sustainable Development 375

• UN Conference on Environment and Development (1992)


• General Assembly Special Session on the Environment (1997)
• World Summit on Sustainable Development (2002)
• UN Conference on Sustainable Development (2012)
• Paris Climate Conference (2015)
Several of these conferences have resulted in significant international
agreements, often termed “conventions,” that provide frameworks for coun-
tries to cooperate in protecting species, habitats, ecosystem processes, and
genetic variation. Treaties that are negotiated at international conferences
come into force when they are ratified by a certain number of countries and
then implemented and enforced at the national level (Figure 11.6). Those
that specifically protect species, such as the Convention on International
Trade in Endangered Species (CITES), were discussed in Chapter 6. Next
we elaborate on the most important conferences and other international
agreements that impact habitat use more generally, especially in the context
of sustainable development.

200
CITES
World Heritage
Number of parties to agreement

CBD
150 UNFCCC
Ramsar
Convention comes
into force
100

50

0
1970 1975 1980 1985 1990 1995 2000 2005 2010 2015
Year
Figure 11.6 Major multinational environmental agreements (MEAs) are ne-
gotiated and then ratified by the governments of individual countries, which be-
come “parties,” or participants, in the provisions of the agreements or treaties. A
treaty comes into force (i.e., countries begin to follow the provisions of the treaty)
when it has been signed by a certain number of countries (indicated by a dot).
The plot lines show the numbers of countries that have ratified various treaties
that provide for biodiversity protection by protecting habitat (the Ramsar Con-
vention on Wetlands of International Importance, the World Heritage Conven-
tion/WHC concerning the Protection of the World Cultural and Natural Heritage),
species (the Convention on International Trade in Endangered Species/CITES,
the Convention on Biological Diversity/CBD), and the environment (the United
Nations Framework Convention on Climate Change/UNFCCC). (After WRI 2003,
with updates from MEA websites.)
376 Chapter 11

The United Nations Conference on Environment and Development


(UNCED), held for 12 days in June 1992 in Rio de Janeiro, Brazil, was one
of the most significant steps toward adopting a global approach to sound
environmental management. Known unofficially as the Earth Summit or
the Rio Summit, the conference brought together representatives from 178
countries, including heads of state, leaders of the United Nations, and indi-
viduals from major conservation organizations and other groups represent-
ing religions and indigenous peoples. Their purpose was to discuss ways
of combining increased protection of the environment with sustainable
economic development in less-wealthy countries (United Nations 1993a,b).
In addition to initiating many new projects, conference participants
discussed, and most countries eventually signed, four major documents:
1. The Rio Declaration. This nonbinding declaration provides general prin-
ciples to guide the actions of both wealthy and poor nations on issues
of the environment and development. The right of nations to use their
own resources for economic and social development is recognized, as
long as the environments of other nations are not harmed in the pro-
cess. The declaration affirms the “polluter pays” principle, in which
companies and governments take financial responsibility for the envi-
ronmental damage that they cause.
2. The United Nations Framework Convention on Climate Change (UNFCCC).
Almost universally ratified (194 signatories), this agreement requires
industrialized countries to reduce their emissions of carbon dioxide
(CO2) and other greenhouse gases and to make regular reports to the
United Nations on their progress. While specific emission limits were
not decided on, the convention states that greenhouse gases should
be stabilized at levels that will not interfere with the Earth’s climate.
3. The Convention on Biological Diversity (CBD). This convention has three
objectives: (1) protecting the various components of biodiversity, (2) using
the components sustainably, and (3) sharing the benefits of new prod-
ucts that are made with the genetic resources of wild and domestic
species (www.cbd.int). Developing international laws for intellectual
property rights that fairly share the financial benefits of biodiversity
among countries, biotechnology companies, and local people is proving
to be a major challenge to the convention. Because of concerns about
the use or misuse of biological materials, certain developing countries
have established highly restrictive procedures for granting permits
to scientists who want to collect biological samples for their research
(Watanabe 2015). In other cases, new research facilities have been built
in developing countries and local people have been trained in scientific
procedures so that biological samples do not have to be exported. In
2010, the participants in the CBD developed a list of goals to achieve
sustainability, called the Aichi Biodiversity Targets (www.cbd.int/sp/
targets). The main goal was to slow or stop the loss of biodiversity by
reducing the impact of human activities. This was to be achieved in
The Challenges of Sustainable Development 377

part by changing governmental policies and increasing the percent-


age of land and ocean under protection; targets are far from being met
(Butchart et al. 2015) (Figure 11.7).

(A)
Costa Rica
60
Planning units needing protection (%)

Haiti

40
Burundi

needing protection (%)


Biodiversity Indonesia

Planning units

High
target India USA
20
China
Russia Canada

Low
Niger Germany Low High
0 GDP (per capita)
2 3 4 5
Per capita GDP (Log10)

(B)

Equator

Figure 11.7 The proportion of 30 × 30 km conservation-planning units need-


ing protection (in addition to existing protected areas). (A) The dotted line on
the graph indicates the Aichi Biodiversity Target of 17%. Wealthier countries (i.e.,
those with a higher per capita gross domestic product [GDP] such as Canada and
Germany) were more likely to have reached this target than developing countries
(such as Haiti and Costa Rica). (B) The greatest number of units needing addi-
tional protection (as indicated in red and pink) are concentrated in tropical areas.
Colors in the graph correspond to those in the map. (After Butchart et al. 2015.)
378 Chapter 11

4. Agenda 21. This 800-page document is an innovative attempt to com-


prehensively describe the policies governments need to implement for
environmentally sound development. Agenda 21 links the environ-
ment with other development issues, which are most often considered
separately, such as child welfare, poverty, gender issues, technology
transfer, and the unequal division of wealth. Plans of action address
the problems of atmospheric, terrestrial, and aquatic pollution; land
degradation and desertification; mountain development; unsustain-
able agriculture and rural development; and deforestation. Financial,
institutional, technological, legal, and educational mechanisms that
governments can use to implement the action plans are also described.
After the UNFCC, international agreements to reduce global greenhouse
gas emissions to below-1990 levels resumed with the Kyoto Protocol in
December 1997. The agreement was ratified in 2004 under the UNFCCC
(see item 2), and many countries have established policies that have reduced
their emissions of greenhouse gases, primarily CO2. There were additional
talks in Bali in December 2007, in Copenhagen in December 2009, in Warsaw
in November 2013, and in Paris in December 2015. The Paris Accord is the
strongest international agreement on climate change yet, requiring action
from all 195 signatory nations. Previous agreements did not require reduc-
tions from China, which has become the greatest contributor of greenhouse
gasses. Goals of the Paris Accord include taking actions to prevent global
temperature from increasing more than 2°C above pre-industrial levels.
The National Climate Action Plans already shared by 186 nations alone are
expected to cut global emissions in half (Davenport 2015).
The World Summit on Sustainable Development, held in Johannesburg,
South Africa, in 2002, emphasized achieving the social and economic goals
of sustainability. This shift in focus from the Rio Summit highlights a
significant, ongoing debate over whether the emphasis in conservation
should be to promote sustainable use of natural resources for the benefit
of poor people or to protect natural areas and biodiversity (Roe et al. 2013).
The UN Conference on Sustainable Development (unofficially the Rio+20),
held in 2012, linked biodiversity conserva-
tion to sustainable development and climate
change and emphasized the need for market-
based solutions (Carrière et al. 2013). It resulted
in what is referred to as the 2030 Agenda,
which includes 169 targets. A working group
from that conference then developed a set of
17 sustainable development agenda items,
presented at the UN Sustainable Development
Summit in 2015 (Figure 11.8). On January 1,
Figure 11.8 A poster advertising the World Sum- 2016, the 17 goals became officially in force
mit on Sustainable Development. The logo of colored from that meeting, with the ambitious aims to
arrows moving forward reflects the focus on social eliminate poverty, ensure equitability (social
and economic goals over biodiversity goals. justice), and stop climate change.
The Challenges of Sustainable Development 379

International agreements that protect habitat


Seven international conventions protect biodiversity either di-
rectly or through the protection of habitat (Table 11.1) and each Countries can gain
seeks to balance conservation concerns with development needs. international recognition
Those that include provisions for land use and sustainable de- for protected areas
velopment include: (1) the Ramsar Convention on Wetlands, through the Ramsar
(2) the World Heritage Convention (WHC) (or the Convention Convention, the WHC,
Concerning the Protection of the World Cultural and Natural and the Biosphere
Heritage), and (3) the Convention on Biological Diversity (CBD).
Reserves Program.
The Ramsar Convention on Wetlands (www.ramsar.org)
Transfrontier parks in
was established in 1971 to halt the continued destruction of
border areas provide
wetlands, particularly those that support migratory waterfowl,
opportunities for both
and to recognize the ecological, scientific, economic, cultural,
conservation and
and recreational values of wetlands (Figure 11.9). The Ramsar
Convention covers freshwater, estuarine, and coastal marine international cooperation.
habitats and includes 2170 sites with a total area of more than
207 million ha. The 168 countries that have signed the Ramsar

Table 11.1 The Seven Biodiversity-related International Agreements,


with Their Primary Objectives
Convention name and primary objective
Convention on Biological Diversity (CBD): Promotes the
conservation of biological diversity, sustainable use, and equitable
sharing of benefits.

Convention on International Trade in Endangered Species


(CITES): Ensures that trade in animals and plants does not threaten
their survival.

Convention on the Conservation of Migratory Species of Wild


Animals: Provides guidelines for the conservation and sustainable use
of migrating animals throughout their ranges.

Convention on Wetlands (aka the Ramsar Convention): Promotes


the conservation and sustainable use of wetlands and their resources
that contribute to both biodiversity and human well-being.

UNESCO Man and the Biosphere Program: Biosphere Reserves are


internationally recognized areas established to balance human and
biodiversity needs.

International Treaty on Plant Genetic Resources for Food and


Agriculture: Promotes the conservation of plant genetic resources and
the equitable sharing of the benefits that arise from them.

World Heritage Convention (WHC): Mandates the identification and


conservation of the world’s cultural and natural heritage by protecting a
specific list of sites.

Source: After Convention on Biological Diversity, www.cbd.int/brc.


380 Chapter 11

Figure 11.9 The Djoudj National Bird Sanctuary is a Ramsar-listed site


in Senegal noted for millions of waterbirds, such as these great white pelicans
(Pelecanus onocrotalus), which stop here during their annual migration south
from Europe in autumn. (© BSIP/UIG/Getty Images.)

Convention agreed to conserve and protect their wetland resources and


designate at least one wetland site of international significance for conser-
vation purposes. The fourth Ramsar Strategic Plan was launched August
21, 2015, to cover the period from 2016 to 2024.
The goal of the World Heritage Convention is to protect cultural areas
and natural areas of international significance through its World Heritage
Site program (whc.unesco.org). The convention is unique because it em-
phasizes the cultural as well as the biological significance of natural areas
and recognizes that the world community has an obligation to financially
support the sites. Limited funding for World Heritage Sites comes from
the United Nations Foundation, which also supplies technical assistance.
As with the Ramsar Convention, the World Heritage Convention seeks
to give international recognition and support to protected areas that are
established initially by national legislation. The 981 World Heritage Sites
protecting natural areas cover about 142 million ha and include some of the
world’s premier conservation areas: Serengeti National Park in Tanzania,
Sinharaja Forest Reserve in Sri Lanka, Iguaçu Falls in Brazil (Figure 11.10),
Manu National Park in Peru, the Queensland Rain Forest of Australia,
Komodo National Park in Indonesia, and Great Smoky Mountains National
Park in the United States, to name a few.
UNESCO’s World Network of Biosphere Reserves was founded in 1971.
Biosphere reserves are designed to be models that demonstrate the compat-
ibility of conservation efforts and sustainable development for the benefit
of local people, as described in Chapter 8 (see Figure 8.19). A total of 621
biosphere reserves have been created in 117 countries, covering more than
263 million ha and including 47 reserves in the United States, 41 in Mexico,
The Challenges of Sustainable Development 381

Figure 11.10 World Heritage Sites include some of the most revered and
well-known conservation areas in the world. (© Joris Van Ostaeyen/istock.)

41 in Russia, 32 in China, 16 in Bulgaria, and 15 in Germany. The largest


biosphere reserve, located in Greenland, is more than 97 million ha in area.
Along with provisions of the CBD (see the next section, “Funding for
Conservation”), these three conventions establish an overarching consen-
sus regarding the appropriate conservation of protected areas and certain
habitat types. More limited international agreements protect unique eco-
systems and habitats in various regions, including the Western Hemi-
sphere, the Antarctic, the South Pacific, Africa, the Caribbean, and the
European Union. Other international agreements have been ratified to
prevent or limit pollution that poses regional and international threats
to the environment. For example, the Convention on Long-Range Trans-
boundary Air Pollution in the European region recognizes the role that the
long-range movement of air pollution plays in acid rain, lake acidification,
and forest dieback; the Convention for the Protection of the Ozone Layer
was signed in 1985 to regulate and phase out the use of chlorofluorocar-
bons; and the Convention on the Law of the Sea promotes the peaceful use
and conservation of the world’s oceans.
Conservation measures can also potentially contribute to promoting
cooperation between governments. Such is often the case when countries
need to manage areas collectively. In many areas of the world, largely un-
inhabited mountain ranges mark the boundaries between countries. These
areas often are designed as national parks, with each country managing
its own wildlife and ecosystems. As an alternative, countries can establish
transfrontier parks on both sides of boundaries to cooperatively manage
382 Chapter 11

Figure 11.11 The Greater Limpopo


Transfrontier Park, which includes Gonar- ZIMBABWE Mavue
ezhou National Park, Kruger National Park, Clipobila
Limpopo National Park, and several small- Gonarezhou Sav
Mbizi National Park e Ri Zinave
er conservation areas, has the potential ver
National
to unite wildlife management activities Park
in South Africa, Mozambique, and Zimba- Sengwe
bwe. A larger transfrontier conservation Corridor
area (hashed area) will include Zinave
National Park, Banhine National Park, pri-
vate game reserves, and private farms and
ranches. (After www.peaceparks.org.) Banhine
Mapai National Park

Kruger

Lim
Limpopo
National National

pop
Park Park

o
MOZAMBIQUE

Riv
er
Border post

SOUTH
AFRICA

Komatipoort
Africa

whole ecosystems and promote conservation on a large scale (Thondhlana


et al. 2015). An early example of this collaboration was the decision to man-
age Glacier National Park in the United States and Waterton Lakes National
Park in Canada as the Waterton–Glacier International Peace Park. Today,
intensive efforts are being made to link national parks and protected areas
in Zimbabwe, Mozambique, and South Africa into the Greater Limpopo
Transfrontier Park and other, larger management units (Figure 11.11). An
added advantage of this joint management is that the seasonal migratory
routes of large animals will be protected. As another example, the estab-
lishment of the Red Sea Marine Peace Park between Israel and Jordan is
important, not only for conservation, but also for its potential for building
trust in a war-ravaged region.
Marine pollution is another issue of vital concern because of the ex-
tensive areas of international waters not under national control and the
ease with which pollutants released in one area can spread to another
area. Agreements covering marine pollution include the Convention on
Introduction to Conservation Biology 1E Primack/Sher
Sinauer Associates
Morales Studio/SA
Primack_Sher1E_11.11 Date 3-2-16
The Challenges of Sustainable Development 383

the Law of the Sea, the Regional Seas Program of the United Nations En-
vironmental Programme (UNEP), and the Convention on the Prevention
of Marine Pollution by Dumping of Wastes and Other Matter. Regional
agreements cover the northeastern Atlantic Ocean, the Baltic Sea, and
other specific locations, particularly in the North Atlantic region. The
pelagic zone of the open ocean (the area of the ocean far from the shore)
is still largely unexplored and unregulated at this point and is in urgent
need of protection.

Funding for Conservation


In general, there has been movement away from funding conservation
efforts solely with local taxpayer dollars in favor of a variety of other ap-
proaches. Nongovernmental organizations and policies that require indus-
try and development to pay for mitigation have been important sources of
funding within developed countries. However, funding of conservation in
developing countries has been less straightforward. One of the most con-
tentious issues for international conferences and treaties has been decid-
ing how to fund the proposals, particularly the CBD and other programs
related to sustainable development, conservation, and climate change. Over
the past 20 years, international funding for conservation by developed
countries, foundations, and private donors has increased, though not as
much as developing countries and conservation biologists had hoped. In
anticipation of the UN Sustainable Development Summit in September
2015, a group of officials from various nations convened in Ethiopia’s capi-
tal for the third International Conference on Financing for Development
(FFD3). They produced the Addis Ababa Agenda, intended as the global
plan of how to implement and support the post 2015 agenda (Bhattacharya
et al. 2015). However ambitious and well-intentioned, like many agreements
of this type it was immediately attacked for lacking any concrete commit-
ments or “teeth” to increase funding (Barcia 2015).

The World Bank and international nongovernmental


organizations (NGOs)
At the time of the Earth Summit, the cost of conservation programs was
estimated to be about $600 billion per year, of which $125 billion was to
come from developed countries as part of their overseas development as-
sistance (ODA). Because the level of ODA from all countries in the early
1990s totaled approximately $60 billion per year, implementing these con-
ventions would have required a tenfold increase in the aid commitment at
that time. The developed countries did not agree to this increase in fund-
ing, but they offered an alternative: each country would increase its level
of foreign assistance to 0.7% of its gross national product (GNP) by the
year 2000, which would have roughly doubled the ODA from developed
nations. In the past 15 years, the ODA has been increasing steadily, now
66% more than in 2000. However, no schedule was set to meet the tar-
384 Chapter 11

get date, and as of 2014, only a few wealthy northern European countries
had met the GNP target percentage: Luxembourg (1.06%), Sweden (1.09%),
Norway (1.00%), Denmark (0.86%), and the United Kingdom (0.70%). The
United States is the largest contributor by volume at $32.7 billion in 2014,
but as this represents only 0.19% GNP, it is still well below the 0.70% target
(www.oecd.org). In the United States government, funding for international
conservation programs is spread across many departments, including the
Agency for International Development, the National Science Foundation,
the Smithsonian Institution, and the USFWS.
Through the 1980s, international funding for conservation projects was
approximately $200 million per year, but starting in the early 1990s, it in-
creased to $1 billion per year (Figure 11.12). Though this funding increase
was dramatic, it was still not as much as originally promised. Much of the
increase in conservation funding by developed countries has been chan-
neled through the World Bank (www.worldbank.org) and the associated
Global Environment Facility (GEF) (www.thegef.org/gef). The World Bank
is a multilateral development bank established to promote international
trade and economic activity. It is governed mainly by developed countries,
and only a small portion of its activities are related to conservation. The re-
lated organization, the GEF, was established specifically to channel money
from developed countries to conservation and environmental projects in
Biodiversity aid (billions 2000 U.S. dollars)

1.5

1.0

0.5

0
1980 1985 1990 1995 2000 2005 2010
Year

Figure 11.12 International funding for biodiversity projects on an annual


basis for the period 1980–2008. Annual fluctuations in funding are caused by the
tendency to bunch large projects together and by the momentum created by the
Rio Earth Summit in 1992. The total funding over the entire time period is $18.6 bil-
lion, of which 31% is from the World Bank and 28% is from the Global Environment
Facility (GEF). Another 27% of the funding is provided by donor countries, includ-
ing the United States, the Netherlands, and Germany. (After Miller et al. 2013.)
The Challenges of Sustainable Development 385

developing countries, with much of its funding distributed by the World


Bank, other multilateral development banks with a regional focus, and the
United Nations Development Programme.
Over the past 30 years, a total of approximately $18 billion in interna-
tional funding has been provided for almost 10,000 biodiversity projects
in 171 countries (Miller et al. 2013). After the World Bank ($6.8 billion) and
the GEF ($5 billion), the top donors were the United States ($1.3 billion),
the Inter-American Development Bank ($1.1 billion), and the Netherlands
($783 million). Other large sources of project funding come from the United
Nations Development Programme and UNEP. These organizations provide
funds to national governments and conservation organizations for proj-
ects including establishing protected areas, protecting endangered species,
restoring degraded habitats, training conservation staff, developing con-
servation infrastructure, addressing global climate change, and managing
forest, freshwater, and marine resources (Hickey and Pimm 2011).
The enormous impact of such international funding is illustrated by
the joining of the World Bank, a source of funding, with the World Wild-
life Fund (WWF), an international conservation organization, to initiate a
new program, the Forest Alliance, to protect and mange over 100 million
ha of forest in countries around the world (worldwildlife.org) (Bowler et
al. 2012). The World Bank is also one of the leaders in efforts to reduce
CO2 emissions caused by deforestation in tropical countries including In-
donesia. Through its Forest Carbon Partnership Facility, companies and
developed countries can offset their present production of greenhouse
gases by purchasing carbon credits for maintaining these tropical forests.
The World Bank has partnered with the WWF and other large NGOs in
implementing such programs.
International conservation NGOs (also called INGOs) (e.g., the WWF,
Conservation International, BirdLife International, The Nature Conser-
vancy, and the Wildlife Conservation Society) implement conservation
activities directly, often through a carefully articulated set of priorities
and programs (Robinson 2012). These NGOs have also emerged as leading
sources of conservation funding, raising funds from membership dues,
donations from wealthy individuals, sponsorship from corporations, and
grants from foundations and international development banks (Figure
11.13). Although foundations and conservation organizations provide only
a small fraction of the funding for biodiversity projects, they can sometimes
be more flexible and can fund innovative small projects and provide more
intensive management than is typical from projects with government fund-
ing. The growing importance of private funding throughout the world
is illustrated by a recent $260 million donation given by the Gordon and
Betty Moore Foundation in the United States to the NGO Conservation
International.
International NGOs are often active in establishing, strengthening,
and funding both local NGOs and government agencies in the develop-
ing world that run conservation programs (Cohen-Shacham et al. 2015).
386 Chapter 11

Figure 11.13 Over the past 200


four decades, there has been a WWF (World Wildlife Fund)
dramatic increase in the annual SC (Sierra Club)
contributions to many conserva- EDF (Environmental Defense Fund)
150

(millions of U.S. dollars)


tion organizations, as illustrated TNC/100 (The Nature Conservancy)

Annual revenue
by the revenues of four large
NGOs from the United States: The
Nature Conservancy, World Wild- 100
life Fund (WWF), Environmental
Defense Fund, and Sierra Club.
Note that the revenue for The
50
Nature Conservancy should be
multiplied by 100; for 2013, con-
tributions were approximately $5
billion. (After Zaradic et al. 2009, 0
with updates from P. Zaradic.)

81

11
72

78

90

93

99

02

08
75

84

87

96

05
19

20
19

19

19

19

19

20

20
19

19

19

19

20
Year

From the perspective of an INGO such as the WWF, working with local
organizations in developing countries is an effective strategy because it
relies on local knowledge and it trains and supports groups of citizens
within the country, who can then be conservation advocates for years to
come. NGOs are often perceived to be more effective at carrying out con-
servation projects than government departments, but programs initiated
by NGOs may end after a few years when funding runs out, and often
they fail to achieve a lasting effect. Moreover, the income of NGOs can be
quite variable, depending on the state of the economy.

Environmental trust funds


In addition to direct grants and loans for projects provided by the World
Bank and other institutions, another important mechanism used to pro-
vide secure, long-term support for conservation activities in developing
countries is the national environmental fund (NEF). NEFs are typically set
up as conservation trust funds or foundations in which a board of trust-
ees—composed of representatives of the host government, conservation
organizations, and donor agencies—allocates the annual income from an
endowment to support inadequately funded government departments and
nongovernment conservation organizations and activities. NEFs have been
established in over 50 developing countries with funds contributed by
developed countries and by major organizations such as the World Bank,
the GEF, and the WWF.
One important early example of an NEF, the Bhutan Trust Fund for
Environmental Conservation (BTF), was established in 1991 by the govern-
ment of Bhutan in cooperation with the World Bank and the WWF. The BTF
has already received about $44 million (exceeding its goal of $20 million).
The fundIntroduction
providestomore than Biology
Conservation $1 million per year for surveying the rich
1E Primack/Sher
Sinauer Associates
Morales Studio/SA
Primack_Sher1E_11.13 Date 3-2-16 3-14-16
The Challenges of Sustainable Development 387

biological resources of this eastern Himalayan country. NEFs have prolifer-


ated in recent years, with the Latin American and Caribbean Network of
Environmental Funds (RedLAC) alone comprising 13 countries and over
3000 projects supported by an annual budget of over $70 million (www.
redlac.org).

Debt-for-nature swaps
Many countries in the developing world have accumulated huge
international debts that they are unable to repay. As a result, some Government, World Bank,
developing countries have rescheduled their loan payments, uni- and NGO funding for
laterally reduced them, or stopped making them altogether. Be- conservation projects
cause of the low expectation of repayment, the commercial banks has increased in recent
that hold these debts have sometimes sold them at a steep dis- decades. Environmental
count on the international secondary debt market. For example, trust funds and debt-for-
Costa Rican debt has traded for only 14%–18% of its face value. nature swaps provide
In a creative approach, debt from the developing world is additional mechanisms
used as a vehicle for financing projects to protect biodiversity to support conservation
in so-called debt-for-nature swaps (Cassimon et al. 2014). In one activities.
common type of debt-for-nature swap, an NGO in the developed
world (such as Conservation International) buys up the debts
of a developing country; the NGO agrees to forgive the debt in
exchange for the country carrying out a conservation activity. This activity
could involve land acquisition for conservation purposes, park manage-
ment, development of park facilities, conservation education, or sustainable
development projects. In another type of swap, a government of a devel-
oped country that is owed money directly by a developing country may
decide to cancel a certain percentage of the debt if the developing country
will agree to contribute to an NEF or some other conservation project.
Debt-for-nature swaps have converted debt valued at $1.5 billion into
conservation and sustainable development activities in Colombia, Poland,
the Philippines, Madagascar, and a dozen other countries. However, spend-
ing money on conservation programs may divert funds from other neces-
sary domestic programs such as medical care, schools, and agricultural
development. Furthermore, these programs have become less common
because external debt of countries is not generally available at such steep
discounts.

How effective is conservation funding?


Conservation organizations have developed a number of tools to evaluate
the effectiveness of funded projects (Mascia et al. 2014). Projects funded by
the GEF and World Bank so far have received mixed evaluations (www.
thegef.org/gef). On the positive side, increased funding was available for
conservation and biodiversity projects, planning national biodiversity strat-
egies and legislation, identifying and protecting important ecosystems
and habitats, and enhancing staff training. However, the lack of participa-
tion by community groups, local scientists, and government leaders; an
388 Chapter 11

overreliance on foreign consultants; and an elaborate and time-consuming


application procedure were identified as major problems. An additional
problem was the mismatch of short-term funding with the long-term needs
of poor countries.
It must be recognized that many environmental projects supported
by international aid do not provide lasting solutions to the problems be-
cause of failure to deal with the “4 Cs”—concern, contracts, capacity, and
causes. Environmental aid will be effective only when applied to situations
in which both donors and recipients have a genuine concern to solve the
problems (Do key people really want the project to be successful, or do
they just want the money?); when mutually satisfactory and enforceable
contracts for the project can be agreed on (Will the work actually be done
once the money is given out? Will money be siphoned off into private
hands?); where there is the capacity to undertake the project in terms of
institutions, personnel, and infrastructure (Do people have the
skills to do the work, and do they have the necessary resources,
While the recent such as vehicles, research equipment, buildings, and access to
increased funding for the information, to carry out the work?); and when the causes of the
protection of biodiversity problem are addressed (Will the project treat the underlying
is welcome, further
causes of the problem or just provide temporary relief of the
symptoms?). Despite these problems, international funding of
funding is needed to
conservation projects continues. Past experiences are inform-
accomplish the task.
ing new projects, which are more effective, but with the result
that the application and accounting processes can be extremely
cumbersome and time-consuming.
The need for increased funding for biodiversity remains great at the
local, national, and international levels. It is estimated that it will take $3.4–
$4.8 billion annually to reduce the extinction risk of all globally threatened
species by just one International Union for Conservation of Nature (IUCN)
category (e.g., from GS1 to GS2), and that adequately managing and protect-
ing all areas with such species will cost $76 billion annually (McCarthy et al.
2012). While $76 billion is an enormous amount of money, it is dwarfed by
the more than $750 billion spent on the military defense of the United States
in 2014. Similarly, while the conservation funds provided by the United
States government and the World Bank seem large, they are small compared
with the other activities that these and other major international organiza-
tions support. Certainly the world’s priorities could be modestly adjusted
to give more resources to the protection of biodiversity (Rands et al. 2010).
There is also a role to be played by conservation organizations and busi-
nesses working together to market “green products.” Already the Forest
Stewardship Council and similar organizations are certifying wood prod-
ucts from sustainably managed forests and coffee companies are market-
ing shade-grown coffee. If consumers are educated to buy these products
at a somewhat higher price, this could be a strong force in international
conservation efforts.
The Challenges of Sustainable Development 389

Finally, a potentially huge new funding source to protect tropical for-


ests is being adopted as part of international efforts to address climate
change. Because about 20% of global greenhouse gas emissions result from
destruction of the tropical forest, a funding mechanism called Reducing
Emissions from Deforestation and Forest Degradation (REDD) could pay
to protect tropical forests (Munawar et al. 2015). REDD rewards poorer
nations for preserving forests by paying them for the carbon that is stored
in their forests. There are huge concerns about whether this money will
be well spent in protecting forests and reducing poverty in developing
countries or whether it will be diverted to other purposes or cause even
worse deforestation in other places (Phelps et al. 2013). Organizations at
all levels will be involved in designing, implementing, and monitoring
what happens as REDD becomes a reality. These types of new and exist-
ing funding mechanisms will all be needed in coming years as part of
international efforts to protect biodiversity.

Summary
„„Sustainable development is economic needed because species migrate across
development that satisfies the present borders, there is an international trade
and future needs of human society while in biological products, the benefits of
minimizing its impact on biodiversity. biological diversity are of international
Achieving sustainable development is a importance, and the threats to diversity
challenge for conservation biology and are often international in scope and re-
society. quire international cooperation.
„„Legal efforts to protect biodiversity „„The 1992 Rio Summit (also known as the
occur at the local, regional, and national Earth Summit) resulted in four major
levels and regulate activities affecting documents that are the foundation of
both private and public lands. Govern- many programs and subsequent interna-
ments and private land trusts buy land tional sustainability conferences.
for conservation purposes or acquire „„Conservation groups, governments in
CEs and development rights for future developed countries, and the World
protection. Associated laws limit pol- Bank are increasing funding to protect
lution, regulate or ban certain types of biodiversity, especially in developing
development, and set rules for hunting countries. NEFs and debt-for-nature
and other activities—all with the aim of swaps are also used to fund conserva-
preserving biodiversity and protecting tion activities. However, the amount of
human health. money is still inadequate to deal with
„„International agreements and conven- the problems.
tions that protect biological diversity are
390 Chapter 11

For Discussion
1. What are the roles of government agen- conservation biologists and environmen-
cies, private conservation organiza- tal activists can energize and enrich each
tions, businesses, community groups, other in working toward an economically
and individuals in the conservation of and environmentally stable world?
biodiversity? Can they work together, or 3. Do you believe that we will be successful
are their interests necessarily opposed to in reaching the global targets established
each other? by the Earth Summit (for biodiversity) or
2. How can conservation biologists provide the Paris Accord (for global warming)?
links between basic science and a public Why or why not? If they are not reached,
environmental movement? What sugges- was there a value to setting them?
tions can you make for ways in which

Suggested Readings
Ban, N. C., M. Mills, J. Tam, C. C. Hicks, S. Klain, N. Stoeckl, and 5 others.
2013. A social-ecological approach to conservation planning: Embedding
social considerations. Frontiers in Ecology and the Environment 11: 194–202.
Concern for the rights and welfare of local people must be included when
designing systems of protected areas.
Brundtland, G., M. Khalid, S. Agnelli, S. Al-Athel, B. Chidzero, L. Fadika, and
M. M. de Botero. 1987. Our Common Future: Report of the World Commission
on Environment and Development (the “Brundtland Report”). This product
of the United Nations World Commission on Environment and Develop-
ment (WCED) laid the groundwork for the important international sum-
mits and agreements to follow, including the 1992 Earth Summit.
Colglazier, W. 2015. Sustainable development agenda 2030. Science 349: 1048–
1050. The Global Sustainable Development Report can be used as a bridge
between the United Nations Sustainable Development Goals and scientific
communities.
Dernbach, J. C. and F. Cheever. 2015. Sustainable development and its discon-
tents. Transnational Environmental Law 4(2): 247–287. Provides answers to
many of the criticisms of sustainable development and a critique of the so-
called alternatives to sustainability.
Fehr-Duda, H. and E. Fehr. 2016. Sustainability: Game human nature. Nature
530. Sustainable development can be promoted by utilizing particular as-
pects of human nature.
Minteer, B. A. and T. R. Miller. 2011. The new conservation debate: Ethical
foundations, strategic trade-offs, and policy opportunities. Biological Con-
servation 144: 945–947. Part of a special issue of the journal devoted to the
balance between protecting biodiversity and providing opportunities for
rural people.
Muradian, R., M. Arsel, L. Pellegrini, F. Adaman, B. Aguilar, B. Agarwal, and
21 others. 2014. Payments for ecosystem services and the fatal attraction
of win–win solutions. Conservation Letters 6: 274–279. Programs involving
PES are being presented as an outstanding opportunity, but they are un-
likely to be as successful as promised.
The Challenges of Sustainable Development 391

Naeem, S, A. V. Ingram, T. Agardy, P. Barten, G. Bennett, E. Bloomgarden, and


39 others. 2015. Get the science right when paying for nature’s services.
Science 347(6227): 1206–1207. Science needs to play a larger role in deter-
mining PES.
Negrón-Ortiz, N. 2014. Pattern of expenditures for plant conservation under
the Endangered Species Act. Biological Conservation 171: 36–43. Plants re-
ceive far less funding than animals.
Nieto-Romero, M., A. Milcu, J. Leventon, F. Mikulcak, and J. Fischer. 2016. The
role of scenarios in fostering collective action for sustainable development:
Lessons from central Romania. Land Use Policy 50: 156–168. Barriers to (and
solutions for) movement toward sustainable development in rural commu-
nities are explored through interviews with Transylvanians.
Nolte, C., A. Agrawal, K. M. Silvus, and B. S. Soares-Filho. 2013. Governance
regime and location influence avoided deforestation success of protected
areas in the Brazilian Amazon. Proceedings of the National Academy of Sci-
ences USA 110: 4956–5961. Once a protected area is established, a variety of
factors can affect its ability to protect biodiversity.
Reed, S. E., J. A. Hilty, and D. M. Theobald. 2014. Guidelines and incentives
for conservation development in local land-use regulations. Conservation
Biology 28: 258–268. Conservation development involves building hous-
ing in a dense cluster, with the remaining land protected for conservation
purposes.
Selomane, O., B. Reyers, R. Biggs, H. Tallis, and S. Polasky. 2015. Towards in-
tegrated social-ecological sustainability indicators: Exploring the contri-
bution and gaps in existing global data. Ecological Economics 118: 140–146.
The contribution of nature to development is largest for poor rural
communities.
Sorice, M. G., C-O. Oh, T. Gartner, M. Snieckus, R. Johnson, and C. J. Donlan.
2013. Increasing participation in incentive programs for biodiversity con-
servation. Ecological Applications 23: 1146–1155. Innovative programs are
being developed to encourage local participation in conservation projects.
Wunder, S. 2013. When payments for environmental services will work for
conservation. Conservation Letters 6: 230–237. PES are best suited to stable
societies with good legal and financial institutions and may not be suited
to many developing countries.
An Agenda for
12 the Future
Ongoing Problems and Possible
Solutions 394
The Role of Conservation
Biologists 402

The future of species like this rare Penland’s beardtongue


flower (Penstemon penlandii) depends on the work of
conservation scientists collaborating with others.
A
s we have seen throughout this book, the
causes of the rapid, worldwide decline in
biodiversity are no mystery. Ecosystems are
destroyed and species are driven to extinction by hu-
man resource use, which is propelled by the need of
many rural people to survive, by the excessive con-
sumption of resources by affluent people and coun-
tries, and by the desire to make money (Fischer et al.
2012). The destruction may be caused by local people
in the region, people recently arrived from outside
the region, local business interests, large businesses
in urban centers, suburban sprawl into rural areas,
multinational corporations in other countries, military
conflicts, governments, or others. Moreover, people
may also be unaware of, or apathetic toward, the im-
pact of human activities on the natural world. Further-
more, although there have never before been so many
people involved in documenting new species, hun-
dreds of thousands of species are yet to be discovered
and may be lost before they are even named (Costello
et al. 2015).
For conservation policies to work, people at all
levels of society must believe that the conservation of
biodiversity is in their own interest (Manfredo et al.
2016). If conservationists can demonstrate that pro-
tecting biodiversity has more value than destroying
394 Chapter 12

it, people and their governments may become more willing to preserve
biodiversity. This assessment should include not only immediate monetary
value but also less tangible aspects, including existence value, option value,
and intrinsic value (see Chapter 3).

Ongoing Problems and Possible Solutions


There is a consensus among conservation biologists that our efforts to pre-
serve biodiversity face several major problems and that certain changes
must be made to policies and practices (Gustafsson et al. 2015; Sutherland
et al. 2011). We list some of these problems and suggest some solutions
next. Note that for the purposes of this text we have simplified the solu-
tions, leaving out many of the intricacies that must be addressed to provide
comprehensive, real-world answers.

Problem: Protecting biodiversity is difficult when most of the world’s


species remain undescribed by scientists and are not known by the general
public. Furthermore, most ecosystems lack monitoring to determine how
they are changing over time.
Solution: We must train more scientists and enthusiastic
amateurs to identify, classify, and monitor species and ecosys-
tems, and we must increase funding in these areas (Costello
Conservation biologists
et al. 2015). In particular, we need to train more scientists and
can spread the message
establish research institutes in developing countries. Citizen
of conservation and scientists often can play an important role in protecting and
do better science by monitoring biodiversity if they are given some training and
educating members of guidance by conservation biologists (Danielson et al. 2014)
the public and including (Figure 12.1). People who are interested in conservation biol-
them in their projects. ogy should be taught basic skills, such as species identification
These citizens then often and environmental-monitoring techniques. For example, fol-
become advocates for lowing the Deepwater Horizon oil spill in the Gulf of Mexico,
protecting biodiversity. the US Fish and Wildlife Service, Cornell Lab of Ornithology,
and National Audubon Society quickly mobilized thousands
of volunteers and adapted the online bird-monitoring program
eBird to allow volunteers to collect and submit field observations
that enabled scientists and managers to assess damage to birds and other
wildlife and to target cleanup efforts. Other examples are Project Budburst,
which recruits the public to submit phonological data as a way of tracking
the impacts of climate change, and iNaturalist, an international program
with which amateurs can record observations of plants and wildlife. iNatu-
ralist also connects these enthusiastic amateurs with scientists who can
use their skills to collect data. Such citizen-science programs are becoming
increasingly common and valuable to conservation (Sullivan et al. 2014).
Other types of citizen-science programs are those based in conserva-
tion education. These programs provide specific information to particular
audiences, such as schoolchildren, senior citizens, or rural people living
An Agenda for the Future 395

First adult sightings


Jan–Feb
March 1–March 14
March 15–March 28
March 29–April 11
April 12–April 25
April 26–May 9
May 10–May 23
May 24–June 6
June 7–June 20
After June 20

Figure 12.1 Each spring, thousands of citizen observers, especially stu-


dents, contribute their observations to the Journey North website, which tracks
the migration of the monarch butterfly. This map presents the observations of
the first appearance of monarch butterflies across North America in spring 2009.
The butterflies primarily overwinter in central Mexico and start migrating into
the southern United States in March, arriving in the northern United States and
southern Canada in late May and June. There are secondary overwintering sites
in California, Florida, and Texas. Efforts by citizen observers can be used to
detect the earlier migration of butterflies in response to a warming climate and
a decline of butterfly numbers as a result of habitat damage in the Mexican over-
wintering grounds. (After www.learner.org/jnorth; inset by David McIntyre.)

near national parks, and can help promote conservation-oriented behaviors


(Şekercioğlu 2012). In these types of projects, volunteers not only observe
how to conduct science and conservation, but actually perform science
and conservation. In this way they can experience a sense of service and
ownership of the project and become advocates. In some cases, this work
translates into lasting advances in conservation and can inspire students
to pursue careers in the field. Many conservation biologists and ecologists
attribute their passion for the field to a transformative outdoor experience
such as those provided by these programs.
Other education approaches have been developed that improve people’s
understanding of conservation, including the creative use of smart phones,
web resources, digital photography, and other advanced technology. In-
formation on biodiversity must be made more accessible; this may be ac-
396 Chapter 12

complished in part through two new sites, the Encyclopedia of Life (eol.
org) and the Tree of Life Web Project (tolweb.org), which serve as central
clearinghouses for data. The question of how to improve the science, con-
servation, and education outcomes of citizen science and other educational
techniques is an active area of research.

Problem: Many conservation issues are global in scope and involve


many countries, making it important but difficult to coordinate conserva-
tion actions. Furthermore, international agreements are extremely difficult
to enforce, making noncompliance an issue.
Solution: We must focus on and build upon successful international
efforts. Countries are increasingly willing to discuss international conser-
vation issues, as shown by the 2015 Paris Accord (see Chapter 11). Nations
are also becoming more willing to sign and implement treaties such as the
Convention on Biological Diversity, the United Nations Framework Con-
vention on Climate Change, and the Convention on International Trade in
Endangered Species. International conservation efforts are expanding, and
further participation in these activities by conservation biologists and the
general public should be encouraged. Another positive development is the
trend toward establishing transfrontier parks that straddle national borders;
these parks benefit wildlife and encourage cooperation between countries.
Citizens and governments of developed countries must also become
aware that they bear a direct responsibility for destroying biodiversity
through their overconsumption of the world’s resources and the specific
products they purchase (Sayre et al. 2013) (Figure 12.2A). To encourage the
enactment of legislation, conservation professionals must demonstrate how
changes in the actions and lifestyles of individuals on the local level, and
those of cities and nations on a larger scale, can exert positive influences
far beyond the immediate community. Improvements in understanding
conservation psychology will help in the design of programs to encourage
conservation-minded behaviors.

Problem: Developing countries often want to protect their


biodiversity but face great pressure to develop their natural re-
sources to generate needed income.
Solution: Conservation organizations, zoos, aquariums, bo-
Conservation biologists tanical gardens, natural history museums, governments, and
need to support even businesses in developed countries, as well as international
approaches that provide organizations such as the United Nations and the World Bank,
benefits for people and should continue to provide technical and financial support to
protect biodiversity. developing countries for conservation activities, in particular
One approach is to establishing and maintaining protected areas. It is important
compensate landowners for these organizations to support the training of conservation
and local people for the biologists in developing countries so local people can become
ecosystem services that advocates for biodiversity within their own country (Paknia
their land provides. et al. 2015). Support from developed nations should continue
until countries can protect biodiversity with their own re-
An Agenda for the Future 397

(A) (B) Figure 12.2 (A) A farmer in


Nicaragua grows and processes
sustainably produced coffee
berries in hope of receiving a
higher price for the crop than
for conventionally produced
coffee berries. (B) Sustainably
produced and certified coffee
is available for purchase in an
increasing number of stores and
has a recognizable Fair Trade
logo on the package. (A, © Janet
Jarman/Corbis; B, courtesy of
David McIntyre.)

sources and personnel. Requiring support from developed nations until


that time is fair and reasonable since they have the funds to support these
parks, drive much of the degradation of biodiversity through consumption,
benefit from the exploitation of natural resources in these countries, and
make use of biological resources in their agriculture, industry, research
programs, zoos, aquariums, botanical gardens, and educational systems.
Solutions for economic and social problems in developing countries must
be addressed simultaneously, particularly those relating to reducing pov-
erty and ending armed conflicts (see Chapter 11). A variety of financial
mechanisms exist to achieve these goals, including direct grants, payments
for ecosystem services, debt-for-nature swaps, and trust funds. Individual
citizens in developed countries can donate money and participate in or-
ganizations and programs that further advance these conservation goals.

Problem: Economic analyses often paint a falsely encouraging picture


of development projects that damage the environment. Economic deci-
sion making often fails to include and evaluate ecosystem services and
intrinsic values.
Solution: Cost–benefit analyses must evaluate development projects
comprehensively by comparing potential project benefits with the full
range of costs, including environmental and human costs such as soil
erosion, air and water pollution, greenhouse gas production, declines in
the availability of natural products and other ecosystem services, and the
loss of places for people to live (Maron et al. 2013; Pilgrim et al. 2013). Local
communities and the general public should be presented with all available
information and asked to provide input into the decision process. We must
adopt the “polluter pays” principle, in which industries, governments, and
individual citizens pay for cleaning up the environmental damage their
398 Chapter 12

activities have caused (Szlávik and Füle 2009; Zhu and Zhao 2015). And
financial subsidies to industries that damage the environment—such as
the pesticide, fertilizer, transportation, petrochemical, logging, fishing,
and tobacco industries—should end, particularly subsidies to industries
that damage human health as well (see Chapter 3). Funds from these so-
called perverse subsidies should be redirected to activities that enhance
the environment and human well-being, especially to people whose lands
provide ecosystem services to the public.

Problem: Disadvantaged people simply trying to survive are frequently


blamed for the destruction of the world’s biodiversity.
Solution: In places where local actions lead to biodiversity losses, con-
servationists can help recruit development and humanitarian organiza-
tions that have the skills to assist local people in organizing and developing
sustainable economic activities that do not damage biodiversity (Kahler et
al. 2013). Conservation biologists and conservation organizations increas-
ingly participate in programs for poor, rural areas that promote smaller
families, establish reliable food supplies, provide training in economically
useful skills (Moro et al. 2013), and promote the benefits and pleasures of
positive experiences as opposed to physical possessions. These programs
should be closely linked to efforts aimed at recognizing basic human rights,
especially ownership of the land and the right to a healthy environment,
that is, social justice movements (see Chapter 11).
Conservation organizations and businesses should work together to mar-
ket green products produced by rural communities, with some of the profits
shared with those communities. Already the Forest Stewardship Council
and similar organizations certify wood products derived from sustainably
managed forests, coffee companies market shade-grown coffee (Edwards
and Laurance 2012), and aquariums and ocean conservation organizations
have developed lists of seafood that is harvested sustainably. Products meet-
ing environmental, labor, and developmental standards can be Fair Trade
Certified (Figure 12.2B). As of 2016, over 1210 organizations in 74 develop-
ing countries and representing over 1.5 million farmers and workers were
certified to sell fair trade products, including bananas, tea, cotton, and fresh
fruit (www.fairtrade.net). If consumers choose to buy these certified products
instead of noncertified products, despite the fact that they are slightly more
expensive, their purchases can strengthen local and international conserva-
tion efforts and provide tangible benefits to disadvantaged people in rural
areas. However, the source of these products must be carefully investigated
to ensure that their manufacturers have used truly sustainable production
and are not simply using the certification label for marketing purposes
(Christian et al. 2013). In some cases, such as some wine from Argentina,
investigators found no meaningful difference between the practices of the
fair trade and conventional wine sectors (Staricco and Ponte 2015).

Problem: Central governments often make decisions about establish-


ing and managing protected areas with little input from the people and
An Agenda for the Future 399

organizations in the affected region. Consequently, local people sometimes


feel alienated from conservation projects and do not support them.
Solution: For a conservation project to succeed, the local inhabitants
must know they will benefit from it and that their involvement is important
(see Chapter 11). To achieve this goal, environmental impact statements,
economic forecasts, and other project information should be made pub-
licly available to encourage open discussion at all stages of the project’s
development and implementation. Local people should be provided with
the assistance they need to understand and evaluate the implications of
the project. Often they will want to protect biodiversity and associated
ecosystem services because they understand that their livelihood, quality
of life, and sometimes, even their survival depend on protecting the natural
environment (Gupta et al. 2015).
Mechanisms should be established to ensure that government agencies,
conservation organizations, and local communities and businesses share
input into the management process. Conservation biologists working in
national parks and other protected areas should engage in meaningful
dialogue to explain the purpose and results of their work to nearby com-
munities and school groups and listen to what local people have to say.
In some cases, regional strategies, such as habitat conservation plans or
community conservation plans, may have to be developed to reconcile the
need for some development (and resulting loss of habitat) with the need to
protect species and ecosystems.

Problem: Revenues, business activities, and scientific research associ-


ated with national parks and other protected areas may not directly benefit
the people living in the surrounding communities.
Solution: People living in local communities often bear the costs, but
do not receive the benefits, of nearby protected areas. Addressing this
problem requires developing mechanisms to benefit the local communi-
ties (Figure 12.3). For example, when possible, local people should be
trained and employed in protected areas as a way of using local knowledge
and providing income. Governments or organizations can also assist local
people in developing businesses related to tourism and other park activities
A portion of park revenues should fund local community projects such as
schools, clinics, roads, cultural activities, sports programs and facilities,
and community businesses—infrastructure that benefits an entire village,
town, or region; this establishes a link between conservation programs and
the improvement of local lives.

Problem: National parks and conservation areas often have inadequate


budgets to pay for conservation activities.
Solution: It is often possible to increase funds for park management
by raising rates for admission, lodging, or meals so that they reflect the
actual cost of maintaining the area. Concessions selling goods and ser-
vices may be required to contribute a percentage of their income to the
park’s operation. Also, zoos and conservation organizations in the de-
400 Chapter 12

Figure 12.3 A women’s cooperative in Puno, Peru, makes and sells hats
featuring the endangered Lake Titicaca frog to sell to tourists. The cooperative
was started by conservation biologists with the Denver Zoo seeking to create
income streams for local people that did not involve harvesting the animals.
(Photograph by Richard Reading.)

veloped world should continue to make direct financial contributions to


conservation efforts in developing countries. For example, members of the
American Zoo and Aquarium Association and their partners participate
in over 3700 in situ conservation projects in 100 countries worldwide (see
Chapter 7).

Problem: Many endangered species and ecosystems occur on private


land or on government land managed for timber production, grazing, min-
ing, and other activities. Timber companies that lease forests and ranchers
who rent rangeland from the government often damage biodiversity and
reduce the productive capacity of the land in the pursuit of short-term
profits. Private landowners often regard endangered species on their land
as a burden because of restrictions on the use of their habitats.
Solution: Laws should be changed so people are only permitted to
obtain leases to harvest trees and use rangelands if they maintain the
health of the ecosystem (Stocks 2005). Governments should also eliminate
perverse tax subsidies that encourage exploiting natural resources and
establish payments for land management, especially on private land, that
enhance conservation. Additionally, landowners should be educated to
protect endangered species and praised publicly for their efforts. It is vital
to develop connections among farmers, ranchers, conservation biologists,
and hunting and fishing groups because biodiversity, wildlife, and the
rural way of life are all intertwined.
An Agenda for the Future 401

We must also seek ways to integrate biodiversity considerations into


plans, policies, and practices that depend on, or in some way relate to,
the environment, on both the local and the international scales (Redford
et al. 2015). To be effective, this requires communication between those
who want the ecosystem services and those who are more concerned with
the ecosystem itself (QUINTESSENCE Consortium 2016). This approach
has been called mainstreaming biodiversity, and it is frequently used in
the United Nations and other international treaty bodies. Article 10(a) of
the Convention on Biological Diversity (CBD) calls for integrating “as far
as possible and as appropriate, the conservation and sustainable use of
biological resources into national decision-making,” and Decision VII/14
adopted the CBD Addis Ababa Principles and Guidelines for the Sustain-
able Use of Biodiversity (AAPG) in 2004. Spurred by such agreements and
other, more local, motivators, governments are increasingly incorporating
biodiversity into formal laws and standard practices.

Problem: In many countries, governments are inefficient and bound


by excessive regulation. Consequently, they are often slow and ineffective
at protecting endangered species and ecosystems.
Solution: Local nongovernmental organizations (NGOs) and citizen
groups often represent the most efficacious agents for promoting conserva-
tion, especially if they work in concert with government policies, such as
economic incentives for conservation (G. Wright et al. 2015). Accordingly,
these citizen groups should be encouraged and supported politically, sci-
entifically, and financially. Conservation biologists must educate citizens
about local environmental issues and encourage them to take action when
necessary. Building the capacity of universities, the national media, and
NGOs to evaluate, propose, and implement conservation-related policies
also effectively encourages national-level action. Individuals, organiza-
tions, and businesses should start new foundations to financially support
conservation efforts. One of the most important trends in conservation
funding and policy is the increased strength of international NGOs, such
as the World Wide Fund for Nature (with about 5 million members) and the
Royal Society for the Protection of Birds. The number of NGOs has risen
dramatically in past decades, and they can substantially influence local
conservation programs and environmental policy at the local, national,
and international levels (WRI 2003).

Problem: Many businesses, banks, and governments remain uninter-


ested in, and unresponsive to, conservation issues.
Solution: Leaders may become more willing to support conservation
efforts if they receive additional information about the benefits of more
sustainable practices or perceive strong public support for conservation
initiatives (Robinson 2012). In countries with fairly open societies, lobbying
and similar efforts may be effective in changing the policies of unrespon-
sive institutions because politicians and other officials generally want to
402 Chapter 12

avoid bad publicity. Petitions, rallies, letter-writing campaigns,


Conservation biologists and economic boycotts all have their place if requests for change
often work with are ignored. In many situations, radical environmental groups
environmental activists such as Greenpeace and Earth First! dominate media attention
and nongovernmental with dramatic, publicity-grabbing actions, while mainstream
organizations to protect conservation organizations follow after them to negotiate a
biological diversity. This
compromise. In closed societies, focusing on identifying and
educating key leaders is usually a better strategy because of the
often means finding a
dangers faced by any opponents of the government. A better un-
balance between doing
derstanding of the diverse values that different cultures attribute
the best possible science
to biodiversity also can help in promoting sustainable practices.
and political advocacy.

The Role of Conservation Biologists


The problems and solutions outlined here underscore the impor-
tance of conservation biologists, who are among the primary participants
in this arena. Conservation biology differs from many other scientific dis-
ciplines in that it plays an active role in preserving biodiversity in all its
forms: species, genetic variability, biological communities, and ecosystem
functions. As a result, conservation biologists must balance political advo-
cacy with scientific credibility, maintaining the greatest possible objectiv-
ity in their scientific research (Horton et al. 2016). Members of the diverse
disciplines that contribute to conservation biology share the common goal
of protecting biodiversity in practice, rather than simply investigating and
talking about it (Laurance et al. 2013). However, conservationists must
work together to provide practical solutions to address real-world situa-
tions (Redpath et al. 2013).
Working together is easier said than done, and there is concern among
some conservation biologists that disagreements between saving biodiver-
sity for its intrinsic value and biodiversity’s instrumental value (i.e., use
values; see Chapter 3) pose a danger for the future of conservation (Tallis
et al. 2014). In a 2014 issue of Nature, Heather Tallis, Jane Lubchenco, and
238 cosignatories petitioned for cooperation between these camps and sug-
gested that the inclusion of women conservation biologists plays a critical
role in achieving cooperation. Certainly, the inclusion of all bright minds
and good ideas is necessary for the continued success of the discipline.

Challenges for conservation biologists


Decisions about park management and species protection increasingly
incorporate the ideas and theories of conservation biology (e.g., Natural
Resource Stewardship and Science 2015). The need for maintaining large
parks and protecting large populations of endangered species has received
widespread attention in both academic and popular literature. The vulner-
ability of small populations to local extinction, even when they are being
carefully protected and managed, and the alarming rates of species ex-
tinction and destruction of unique ecosystems worldwide have also been
An Agenda for the Future 403

highly publicized. As a result of this publicity, the need to protect biodi-


versity has entered the political debate and been targeted as a priority for
government conservation programs. At the same time, botanical gardens,
museums, nature centers, zoos, national parks, and aquariums are reori-
enting their programs to meet the challenges of protecting biodiversity
and embracing the concepts of Earth stewardship (Sayre et al. 2013). The
sense of urgency is heightened by the recognition that many endangered
species living in cold climates, such as polar bears and penguins, face
immediate threats due to a warming climate and the melting of sea ice.
What is ultimately required, however, is including the principles of con-
servation biology into the broader domestic policy arena and economic-
planning process (Halpern et al. 2013). Incorporating conservation biology
into economic policy and reprioritizing domestic policy goals will require
substantial public education and political effort.

Achieving the agenda


Successfully meeting conservation challenges requires that conservation
biologists take on several active roles. Conservationists must become more
effective educators, leaders, and motivators in the public forum as well as in
the classroom (Figure 12.4). Conservation biologists need to educate as
broad a range of people as possible about the problems that stem from a
loss of biodiversity, convey a positive message about what needs to be done,
and then give examples of some successes (Bickford et al. 2012; Laurance
et al. 2013). The Society for Conservation Biology has made disseminat-
ing knowledge the first item in its new code of ethics. Conservation biolo-
gists must provide actionable research and leadership that counters the
pessimism and passivity so frequently encountered in modern society

Conservation Organize…
biologists Other scientists
Motivate…

General public, Manage… Are politically


especially children active and affect…

Use results
to educate…
Conservation
Public policy
projects Translate results
into…

Figure 12.4 Conservation biologists must be active in various ways to


achieve the goals of conservation biology and to protect biodiversity. Not every
conservation biologist can be active in each role, but all of the roles are important.
404 Chapter 12

(Carlson 2013). The International Union for the Conservation of Nature


demonstrates this approach with their campaign emphasizing “Love. Not
Loss,” which attempts to capture people’s positive feelings toward nature
rather than simply convey dry facts about nature’s demise (www.iucn.org).
Many groups, including anglers, hunters, bird-watchers, hikers, religious
groups, and artists, may help conservation efforts once they become aware
of the issues or recognize that their self-interest or emotional well-being
depends on conservation (Granek et al. 2008). Some evidence suggests that
such awareness is increasing worldwide, even in some of the developing
countries, such as India (Varma et al. 2015).
Conservation biologists often teach college students and write technical
papers addressing conservation issues, but these actions only reach limited
audiences: remember that only a few dozen or at most few thousand people
read most scientific papers. In contrast, millions of adults watch nature pro-
grams on television, especially those produced by the Public Broadcasting
Service and the British Broadcasting Corporation, and tens of millions of
children watch television programs such as Go, Diego, Go and Wild Kratts,
and movies such as Avatar, Wall-E, and The Lord of the Rings, which have
powerful conservation themes. An especially powerful example of the role
of media was the release of Blackfish in 2013, a movie that highlighted the
treatment of captive killer whales (Orcinus orca), including the questionable
practice of how they are captured in the wild. The filmmakers worked with
scientists and rigorously researched all the information presented. The huge
following of the film resulted in legislation proposed in the New York and
California legislatures in 2014 to protect the whales (A. J. Wright et al. 2015),
and Congress unanimously passed an amendment to the US Agriculture
Appropriations Act to update regulations concerning captive marine mam-
mals (Charky 2014). Moreover, on March 16, 2016, SeaWorld announced that
the killer whales currently in its care will be their last (CNN 2016).
Conservation biologists must reach a wider range of people by speak-
ing in villages, towns, cities, elementary and secondary schools, parks,
neighborhood gatherings, and religious organizations (Swanson et al.
2008). Conservationists should also work to more widely incorporate the
themes of conservation into public discussions. Conservation biologists
must spend more of their time writing articles and editorials for news-
papers, magazines, and blogs, as well as effectively speaking on radio,
television, and other mass media, in ways the public can understand (Ja-
cobson 2009). They must make a special effort to talk to children’s groups
and to write versions of their work that children can read. Hundreds of
millions of people visit zoos, aquariums, natural history museums, and
botanical gardens, making these places prime venues for communicating
conservation messages to the public. Conservation biologists must continue
to seek out creative ways to reach wider audiences and avoid repeatedly
“preaching to the converted.” Potential constituents are becoming more
technologically savvy, more diverse, speaking different languages, and
viewing wildlife through different cultural lenses. These changes require
that conservation biologists use new approaches and social media, such
An Agenda for the Future 405

as YouTube, Facebook, Twitter, and blogs, for effective outreach and com-
munication (Jacobson et al. 2014). Data from these approaches can even be
used in conjunction with surveys to evaluate how effective scientists are
in conveying what they intend to convey and whether they are reaching
their intended audiences (Bombaci et al. 2016) (Figure 12.5).
The efforts of Merlin Tuttle and Bat Conservation International (BCI)
illustrate how public attitudes toward even unpopular species can change.
BCI has campaigned throughout the United States and the world to educate

45

40
Intended audience
35
Percentage of audience categories

Audience reached
via social media
30

25

20

15

10

0
n

s
s
ia

er
or
ts

or rs
ia
s

ic

ia
s d

al
ie
io
O

is
m

er an

at o
bl
ed

th
ar
ct

rn
et
G

at

tr ct
nt

s
de

pu

se

O
qu
M

ci
N

ou
uc
ak nt

is re
ie

so
ca

/a

in di
e

sc

ed

cj
at
ym m

th
A

al

os

m m
iv

ifi
h
lic ern

on
of
rc

–1

Pr

ad ra
Zo
nt
sa

si
po ov

rs

& rog
ie
K

es
Re

be

Sc
G

P
of
em

Pr
M

Figure 12.5 During the 2013 International Congress for Conservation Biology,
efficacy of communication was evaluated by comparing the audience intended by
the presenters (as determined via survey) with the audience reached via Twitter,
a social media platform. This graph shows the percentage for each of 13 audience
categories. “Audience reached” was determined by counting the senders and
recipients of 700 live “tweets” during the conference and 1711 “retweets” (forward-
ing of a Twitter message to another set of followers). The mean number of follow-
ers for tweeters and retweeters was 2404, with 40% having over 10,000 followers
across 40 countries; thus, considerably more than the 1500 attendees to the con-
ference were reached. Although academics were an intended audience and were,
in fact, the greatest proportion of people who were reached, the remainder of the
presenters’ intended audience differed from those who were reached via tweets;
the media and program directors were a greater proportion of their audience than
government officials, despite the opposite intent of the presenters. However, these
results suggest that using social media to create media interest may be a good
way for scientists to reach the general public. (After Bombaci et al. 2016.)
406 Chapter 12

people on the importance of bats in ecosystem health, emphasizing their


roles as insect eaters, pollinators, and seed dispersers. A valuable part of
this effort has involved producing bat photographs and films of exceptional
beauty. In Austin, Texas, Tuttle intervened when citizens petitioned the
city government to exterminate the hundreds of thousands of Brazilian
free-tailed bats (Tadarida brasiliensis) that live under a downtown bridge.
He and his colleagues convinced people that the bats are both fun to watch
and critical in controlling insect populations that feed on important crop
plants and bite people. The situation has changed so drastically that the
government now protects these bats as a matter of civic pride and practical
pest control, and hundreds of citizens and tourists gather every night to
watch the swarm of bats emerge from under the bridge on their nightly
foraging forays (Figure 12.6).
Conservation biologists must also become politically active leaders to in-
fluence public policy and new laws (Sale 2015). Scientists can, and should,
participate in politics without sacrificing professional credibility, in part by
practicing transparency and being honest to themselves and others about
the impact of personal values and funding sources (Garrard et al. 2015). In-
volvement in the political process allows conservation biologists to influence
the passage of new laws that support the preservation of biodiversity and to
argue against legislation that would prove harmful to species or ecosystems.

Figure 12.6 Citizens and tourists gather in the evening to watch Brazilian
free-tailed bats emerge from their roosts beneath the Congress Avenue bridge
in Austin, Texas. (Photograph © Merlin D. Tuttle, Bat Conservation International,
www.batcon.org)
An Agenda for the Future 407

An important first step in this process is joining conservation organizations


or mainstream political parties to strengthen them by working within these
groups and learning more about the issues. It is important to note that
there is also room for people who prefer to work by themselves. Difficul-
ties in getting the US Congress to reauthorize the Endangered Species Act
and to ratify the CBD and the United Nations Framework Convention on
Climate Change illustrate the need for greater political activism on the part
of scientists who understand the implications of failing to take immediate
action. Though much of the political process is time-consuming and tedious,
it often represents the only way to accomplish major conservation goals,
such as acquiring new land for reserves or preventing the overexploitation
of old-growth forests. Conservation biologists must master the languages
and methods of legal and political processes and form effective alliances
with environmental lawyers, citizen groups, and politicians.
An example of the effectiveness of alliances between conservation bi-
ologists and other groups is the recent expansion of the Ikh Nart Nature
Reserve in the Dornogobi province of Mongolia. Populations of endangered
cinerous vultures (Aegypius monachus), argali sheep (Ovis ammon), Siberian
ibex (Capra sibirica), and many other species of animals and plants have
been monitored since 2001 by a partnership of conservation biologists
at the Denver Zoo, Mongolian National University, Mongolian National
Education University, University of Vermont, and Mongolian Academy of
Sciences (IkhNart.com). The scientists communicated the findings of their
research to the local government, which responded by doubling the size
of the park to about 130,000 ha and to restore some areas that had been
badly degraded (Reading 2015).
A key role for conservation biologists to assume is that of translators;
that is, scientists who can take the data and results of conservation science
and translate them into legislation, public policy, and management ac-
tions (Courchamp et al. 2015). To be effective, conservation biologists must
communicate the relevance of their research and demonstrate that their
findings are unbiased, while respecting the values and concerns of most
or all stakeholders. Conservation scientists have to be aware of the range
of issues that may affect their programs and that their programs might
affect, and they should be able to speak to general audiences in terms that
they can understand. Conservation biologists must take the lead, as their
expertise is needed.
Conservation biologists should become motivators, convincing a broad
range of people to support conservation efforts and even assisting in for-
mal marketing campaigns to increase awareness (A. J. Wright et al. 2015).
Psychological research suggests that successful campaigns to motivate
people to take conservation actions must include efforts to provide: (1)
information, (2) a sense of belonging, (3) trust, and (4) incentives (Van
Vugt 2009). In general, conservationists must demonstrate to local people
that protecting the environment, not only saves species and ecosystems,
but also improves the long-term health of their families and communi-
408 Chapter 12

ties, their own economic well-being, and their quality of life


The goal of conservation (Morrison 2016). Public discussions, education, and publicity
biology is not just to must be a major part of any such program, and they should
reveal new knowledge, be delivered in ways that instill trust and a sense of belonging
but also to use that and community. In particular, conservationists should devote
knowledge to protect careful attention to convincing business leaders and politicians
biodiversity. Conservation to support conservation efforts. National leaders may be among
biologists must learn to the most difficult people to convince, since they must respond
to diverse interests. However, whether their conversion occurs
demonstrate the practical
due to reason, sentiment, or professional self-interest, once they
application of their work.
join the cause of conservation, these leaders may be in a position
to make major contributions.
Finally, and most importantly, conservation biologists must
become effective managers and practitioners of conservation projects (Blick-
ley et al. 2013). They must be willing to walk on the land and go out on the
water to find out what is really happening, to get dirty, to knock on doors
and talk with local people, and to take risks. Conservation biologists must
learn everything they can about the species and ecosystems they are trying
to protect, and then make that knowledge available to others in a form that
can be readily understood and can affect decision making.
If conservation biologists are willing to put their ideas into practice
and to work with park managers, land-use planners, politicians, and local
people, progress will follow. Getting the right mix of models, new theo-
ries, innovative approaches, and practical examples is necessary for the
discipline to succeed. Once this balance is found, conservation biologists,
working with energized citizens and government officials, will be in a
position to protect the world’s biodiversity during this unprecedented
era of change.

Summary
„„Major problems face us in protecting ries and approaches of their discipline
biodiversity. To address these prob- and actively work with all components
lems, many policies and practices must of society to protect biodiversity and
change. Changes must occur at the local, restore the degraded elements of the
national, and international levels and environment.
will require action on the part of indi- „„To achieve the long-term goals of con-
viduals, conservation organizations, and servation biology, practitioners need to
governments. become involved in conservation educa-
„„Conservation biologists must demon- tion and the political process.
strate the practical value of the theo-
An Agenda for the Future 409

For Discussion
1. Consider a current conflict over conser- Do you think you can make a difference
vation of biodiversity in the news. Does in the world, and if so, in what way?
it fit into one or more of the problem/ 3. Sutherland and colleagues (2009) posed
solution pairs provided in this chapter? 100 questions for conservation biology
What are some possible solutions to the (see Suggested Readings). Provide an-
conflict, given what you now know? swers for the questions you consider to
2. As a result of studying conservation biol- be the most urgent and important.
ogy, have you decided to change your
lifestyle or your level of political activity?

Suggested Readings
Blickley, J. L., K. Deiner, K. Garbach, I. Lacher, M. H. Meed, L. M. Porensky,
and 3 others. 2013. Graduate student’s guide to necessary skills for nonac-
ademic conservation careers. Conservation Biology 27: 24–34. Interpersonal
and project-management skills are needed to conduct practical conserva-
tion work.
Cafaro, P. 2015. Three ways to think about the sixth mass extinction. Biological
Conservation 192: 387–393. How we view the biodiversity crisis will impact
how we respond to it in the future.
Colón-Rivera, R. J., K. Marshall, F. J. Soto-Santiago, D. Ortiz-Torres, and C.
E. Flower. 2013. Moving forward: Fostering the next generation of Earth
stewards in the STEM disciplines. Frontiers in Ecology and the Environment
11: 383–391. Conservation must be made a part of the educational process.
Costello, M. J., B. Vanhoorne, and W. Appeltans. 2015. Conservation of biodi-
versity through taxonomy, data publication, and collaborative infrastruc-
tures. Conservation Biology 29(4): 1094–1099. Collaboration between scien-
tists is essential for protecting biodiversity and should be facilitated.
Danielson, F., K. Pirhofer-Walzi, T. P. Adrian, D. R. Kapijimpanga, N. D. Bur-
gess, P. M. Jensen, and 5 others. 2014. Linking public participation in sci-
entific research to the indicators and needs of international environmental
agreements. Conservation Letters 7: 12–24. Most indicators of environmental
and conservation status can be monitored by trained citizens.
Granek, E. F., E. M. P. Madin, M. A. Brown, W. Figueira, D. S. Cameron, and Z.
Hogan. 2008. Engaging recreational fishers in management and conserva-
tion: Global case studies. Conservation Biology 22: 1125–1134. Recreational
fishers can become strong advocates for conservation.
Jacobson, S. K. 2009. Communication Skills for Conservation Professionals. 2nd ed.
Island Press, Washington, D.C. Researchers have identified practical ways
to increase public support for conservation.
Morrison, S. A. 2016. Designing virtuous socio-ecological cycles for biodiver-
sity conservation. Biological Conservation 195: 9–16. To accomplish suc-
cessful conservation, the public must be educated about the relevance of
biodiversity to their lives and be inspired to act.
410 Chapter 12

Redford, K. H., B. J. Huntley, D. Roe, T. Hammond, M. Zimsky, T. E. Lovejoy,


and 3 others. 2015. Mainstreaming biodiversity: Conservation for the twen-
ty-first century. Frontiers in Ecology and Evolution 3: 137. Considerations of
biodiversity should be incorporated in policies, practices, and principles
for any human activity that affects biodiversity.
Sullivan, B. L., J. L. Aycrigg, J. H. Barry, R. E. Bonney, N. Bruns, C. Cooper and
26 others. 2014. The eBird enterprise: An integrated approach to develop-
ment and application of citizen science. Biological Science 169: 31–40. The
user-friendly approach of the eBird program is engaging tens of thousands
of bird-watchers and thus providing a valuable data source for conserva-
tion biologists.
Sutherland, W. J., W. M. Adams, R. B. Aronson, R. Aveling, T. M. Blackburn, S.
Broad, and 38 others. 2009. One hundred questions of importance to the
conservation of global biological diversity. Conservation Biology 23: 557–567.
Presents 100 scientific questions whose answers would have the greatest
impact on conservation practice and policy.
Tallis, H., Lubchenco, J., Adams, V. M., Adams-Hosking, C., Agostini, V. N.,
and Kovács-Hostyánszki, A. 2014. Working together: A call for inclusive
conservation. Nature 515(7525): 27–28. To be effective, conservation plan-
ning must include women as well as men.
Appendix
Selected Environmental Organizations
and Sources of Information
The following are useful printed resources:
Conservation Directory 2017: The Guide to Worldwide Environmental Organiza-
tions (2016), published by Carrel Press. This directory lists over 4000 local,
national, and international conservation organizations; conservation publica-
tions; and more than 18,000 leaders and officials in the field of conservation.
Pursuing Sustainability: A Guide to the Science and Practice (2016), published
by Princeton University Press. A framework for connecting science with
practice that draws upon the current literature.
The ECO Guide to Careers that Make a Difference: Environmental Work for a
Sustainable World (2004), published by Island Press.

Online searches, especially using Google, provide a powerful way to search


for information concerning people, organizations, places, and topics. The
following are more specialized searchable databases on species and countries:
Encyclopedia of Life IUCN Red List
www.eol.org www.iucnredlist.org
Developing resource for species biology. Information on all plants, fungi and animals
Global Biodiversity Information that have been globally evaluated for
Facility conservation status.
www.gbif.org USDA Plants List
Free and open access to biodiversity data. plants.usda.gov
Data on plants, including distribution, nativity,
and wetland status.

Below is a list of some major conservation organizations and resources:


Association of Zoos and Aquariums (AZA) BirdLife International
8403 Colesville Road, Suite 710 The David Attenborough Building, 1st floor
Silver Spring, MD 20910 USA Pembroke Street
www.aza.org Cambridge, CB2 3QZ, UK
Preservation and propagation of captive www.birdlife.org
wildlife. Determines status, priorities, and conservation
plans for birds throughout the world.
412 Appendix

Center for Plant Conservation European Center for Nature Conservation


15600 San Pasqual Valley Rd. (ECNC)
Escondido, CA 92027, USA P. O. Box 90154
www.centerforplantconservation.org, www. 5000 LG Tilburg, the Netherlands
mobot.org www.ecnc.org
Major center for worldwide plant conservation Provides the scientific expertise that is required
activities. for formulating conservation policy.
Convention on Biological Diversity Secretariat Fauna & Flora International
413 Rue Saint-Jacques, Suite 800 Fauna & Flora International
Montreal, Quebec, H2Y 1N9, Canada The David Attenborough Building
www.cbd.int Pembroke Street
Promotes the goals of the CBD: sustainable Cambridge, CB2 3QZ UK
development, biodiversity conservation, and www.fauna-flora.org
equitable sharing of genetic resources. Long-established international conservation
CITES Secretariat of Wild Fauna and Flora body acting to protect species and ecosystems.
International Environment House Food and Agriculture Organization of the
11 Chemin des Anémones United Nations (FAO)
CH-1219 Châtelaine, Geneva, Switzerland Viale delle Terme di Caracalla
www.cites.org 00513 Rome, Italy
Regulates trade in endangered species. www.fao.org
Conservation International (CI) A UN agency supporting sustainable
2011 Crystal Drive, Suite 500 agriculture, rural development, and resource
Arlington, VA 22202 USA management.
www.conservation.org Friends of the Earth
Active in international conservation efforts and 1101 15th Street NW, 11th floor
developing conservation strategies; home of the Washington, DC 20036 USA
Center for Applied Biodiversity Science. www.foe.org
Earthwatch Institute Attention-grabbing organization working to
114 Western Avenue improve and expand environmental policy.
Boston, MA 02134 USA Global Environment Facility (GEF) Secretariat
www.earthwatch.org 1818 H Street NW, MSN N8-800
Clearinghouse for international conservation Washington, DC 20433 USA
projects in which volunteers can work with www.thegef.org/gef
scientists. Funds international biodiversity and
Environmental Defense Fund (EDF) environmental projects.
1875 Connecticut Ave, NW, Suite 600 Greenpeace International
Washington, DC 20009 USA Ottho Heldringstraat 5
www.edf.org 1006 AZ Amsterdam, the Netherlands
Involved in scientific, legal, and economic www.greenpeace.org/international
issues. Activist organization known for grassroots
Employment Opportunities efforts and dramatic protests against
Various organizations have websites with environmental damage.
environmental and conservation opportunities National Audubon Society
and internships throughout the world: www. 225 Varick Street, 7th floor
webdirectory.com/employment, www.ecojobs. New York, NY 10014 USA
com, to name a few. A publication of interest is www.audubon.org
Careers in the Environment by Mike Fasulo and Involved in wildlife conservation, public
Paul Walker, published by McGraw-Hill. education, research, and political lobbying,
with emphasis on birds.
Organizations and Sources 413

National Council for Science and the Ocean Conservancy


Environment (NCSE) 1300 19th Street NW, 8th floor
1101 17th Street NW, Suite 250 Washington, DC 20036 USA
Washington, DC 20036 USA www.oceanconservancy.org
www.ncseonline.org Focuses on marine wildlife and ocean and
Works to improve the scientific basis for coastal habitats.
environmental decision making; their website Royal Botanic Gardens, Kew
provides extensive environmental information. Richmond, Surrey, TW9 3AB, UK
National Wildlife Federation (NWF) www.kew.org
11100 Wildlife Center Drive The famous Kew Gardens are home to a
Reston, VA 20190 USA leading botanical research institute and an
www.nwf.org enormous plant collection.
Advocates for wildlife conservation. Publishes Sierra Club
the Conservation Directory 2005–2006, as well as 85 Second Street, 2nd floor
the children’s publications Ranger Rick and Your San Francisco, CA 94105 USA
Big Backyard. www.sierraclub.org
Natural Resources Defense Council (NRDC) Leading advocate for the preservation of
40 West 20th Street wilderness and open space.
New York, NY 10011 USA Smithsonian National Zoological Park
www.nrdc.org 3001 Connecticut Avenue NW
Uses legal and scientific methods to monitor Washington, DC 20008 USA
and influence government actions and www.nationalzoo.si.edu
legislation. The National Zoo and the nearby U.S. National
The Nature Conservancy (TNC) Museum of Natural History represent a vast
4245 North Fairfax Drive, Suite 100 resource of literature, biological materials, and
Arlington, VA 22203 USA skilled professionals.
www.nature.org Society for Conservation Biology (SCB)
Emphasizes land preservation. 1133 15th St. NW, Suite 300
NatureServe Washington, DC 20005 USA
4600 North Fairfax Drive, 7th floor www.conbio.org
Arlington, VA 22203 USA Leading scientific society for the field. Develops
www.natureserve.org and publicizes new ideas and scientific results
Maintains databases of endangered species for through the journal Conservation Biology and
North America. annual meetings.
The New York Botanical Garden (NYBG) Student Conservation Association (SCA)
Institute of Economic Botany (IEB) 689 River Road
International Plant Science Center, the New P.O. Box 550
York Botanical Garden Charlestown, NH 03603 USA
2900 Southern Boulevard www.thesca.org
Bronx, NY 10458 USA Places volunteers and interns with
www.nybg.org conservation organizations and public
Conducts research and conservation programs agencies.
involving plants that are useful to people. United Nations Development Programme
Rainforest Action Network (UNDP)
425 Bush Street, Suite 300 1 United Nations Plaza
San Francisco, CA 94108 USA New York, NY 10017 USA
www.ran.org www.undp.org
Works for rain forest conservation and human Funds and coordinates international economic
rights. development activities.
414 Appendix

United Nations Environment Programme Wildlife Conservation Society (WCS)


(UNEP) 2300 Southern Boulevard
United Nations Avenue, Gigiri Bronx, NY 10460 USA
P.O. Box 30552, 00100 www.wcs.org
Nairobi, Kenya Leaders in wildlife conservation and research.
www.unep.org The World Bank
International program of environmental 1818 H Street NW
research and management. Washington, DC 20433 USA
United Nations Environment Programme www.worldbank.org
World Conservation Monitoring Centre (UNEP- Multinational bank involved in economic
WCMC) development; increasingly concerned with
219 Huntingdon Road environmental issues.
Cambridge, CB3 0DL, UK World Conservation Union (IUCN)
www.unep-wcmc.org Rue Mauverney 28
Monitors global wildlife trade, the status of 1196, Gland, Switzerland
endangered species, natural resource use, and www.iucn.org
protected areas. Coordinating body for international
United States Fish and Wildlife Service conservation efforts. Produces directories of
(USFWS) specialists and the Red List of endangered
Department of the Interior species.
1849 C Street NW World Resources Institute (WRI)
Washington, DC 20240 USA 10 G Street NE, Suite 800
www.fws.gov Washington, DC 20002 USA
The leading U.S. government agency concerned www.wri.org
with conservation research and management; Produces environmental, conservation, and
with connections to state governments and development reports.
other government units, including the National
World Wildlife Fund (WWF)
Marine Fisheries Service, the U.S. Forest
1250 24th Street NW
Service, and the Agency for International
Washington, DC 20037 USA
Development, which is active in developing
www.worldwildlife.org, www.wwf.org
nations. The Conservation Directory 2005–2006,
Major conservation organization, with
mentioned previously, shows how these units
branches throughout the world. Active in
are organized.
national park management.
Wetlands International
The Xerces Society
P.O. Box 471
628 NE Broadway, Suite 200
6700 AL Wageningen, the Netherlands
Portland, OR 97232 USA
www.wetlands.org
www.xerces.org
Focus on the conservation and management of
Focuses on the conservation of insects and
wetlands.
other invertebrates.
The Wilderness Society
Zoological Society of London (ZSL)
1615 M Street NW
Outer Circle, Regent’s Park
Washington, DC 20036 USA
London, NW1 4RY UK
www.wilderness.org
www.zsl.org
Devoted to preserving wilderness and wildlife.
Center for worldwide activities to preserve
nature.
Glossary
Numbers in brackets indicate the chapter in which the term is defined.

A the amount people are willing to pay for this goal.


acid rain Rainwater that has become acidic due to Also known as beneficiary value. [3]
air pollution. [4] beta diversity Rate of change of species composi-
adaptive management Implementing a manage- tion along a gradient or transect. [2]
ment plan and monitoring how well it works, then binomial The unique two-part Latin name taxono-
using the results to adjust the management plan. [8] mists bestow on a species, such as Canis lupus (gray
adaptive restoration Using monitoring data to ad- wolf) or Homo sapiens (humans). [2]
just management plans to achieve restoration goals. [10] bioblitz A one-day event in which scientists and
Aichi Biodiversity Targets A list of goals to citizen scientists perform an intensive biological sur-
achieve sustainability and the protection of biodi- vey of a designated area in a short time with the goal
versity developed by the Convention on Biological of documenting all living species in that area. [2]
Diversity (CBD). [11] biocontrol The use of one type of organism, such
Allee effect Inability of a species’ social structure as an insect, to manage another, undesirable, species,
to function once a population of that species falls such as an invasive plant. [10]
below a certain number or density of individuals. [5] biocultural restoration Restoring lost ecological
alleles Different forms of the same gene (e.g., dif- knowledge to people to give them an appreciation of
ferent alleles of the genes for certain blood proteins the natural world. [10]
produce the different blood types found among biodiversity indicators Species or groups of spe-
humans). [2] cies that provide an estimate of the biodiversity in an
alpha diversity The number of different species in area when data on the whole community is unavail-
a community or specific location; species richness. [2] able. Also known as surrogate species. [6]
amenity value Recreational value of biodiversity, biodiversity offsets See compensatory mitiga-
including ecotourism. [3] tion. [10]
arboretum Specialized botanical garden focusing biodiversity The complete range of species, biologi-
on trees and other woody plants. [7] cal communities, and their ecosystem interactions
artificial incubation Conservation strategy that and genetic variation within species. Also known as
involves humans taking care of eggs or newborn biological diversity. [1]
animals. [7] biological community A group of species that oc-
artificial insemination Introduction of sperm into cupies a particular locality. [2]
a receptive female animal by humans; used to increase biological definition of species Among biolo-
the reproductive output particularly of endangered gists, the most generally used of several definitions of
species. [7] “species.” A group of individuals that can potentially
breed among themselves in the wild and that do not
B breed with individuals of other groups. Compare with
background extinction rates The rate of spe- morphological definition of species. [2]
cies loss expected to occur in the absence of human biological diversity See biodiversity. [1]
impact. [5]
biomagnification Process whereby toxins become
beneficiary value See bequest value. [3] more concentrated in animals at higher levels in the
bequest value The benefit people receive by food chain. [4]
preserving a resource or species for their children and
descendants or future generations, and quantified as
416 Glossary

biophilia The postulated predisposition in hu- competition A contest between individuals or


mans to feel an affinity for the diversity of the living groups of animals for resources. Occurs when indi-
world. [1] viduals or a species use a limiting resource in a way
biopiracy Collecting and using biological materi- that prevents others from using it. [2]
als for commercial, scientific, or personal use without conservation banking A system involving devel-
obtaining the necessary permits. [3] opers paying landowners for the preservation of an
bioprospecting Collecting biological materials as endangered species or protected habitat type (or even
part of a search for new products. [3] restoration of a degraded habitat) to compensate for a
bioregional management Management system species or habitat that is destroyed elsewhere. [11]
that focuses on a single large ecosystem or a series conservation biology Scientific discipline that
of linked ecosystems, particularly where they cross draws on diverse fields to carry out research on bio-
political boundaries. [9] diversity, identify threats to biodiversity, and play an
bioremediation The use of an organism to clean up active role in the preservation of biodiversity. [1]
pollutants, such as bacteria that break down the oil in conservation concessions Methods of protect-
an oil spill or wetland plants that take up agricultural ing land whereby a conservation organization pays a
runoff to clean the water [10] government or other landowner to preserve habitat
biosphere reserves Protected areas established as rather than allow an extractive industry to damage the
part of a United Nations program to demonstrate the habitat. [11]
compatibility of biodiversity conservation and sus- conservation corridors Connections between
tainable development to benefit local people. [8] protected areas that allow for dispersal and migra-
biota A region’s plants and animals. [2] tion. Also known as habitat corridors, or movement
corridors. [8]
Bonn Convention Treaty to protect European spe-
cies, particularly migratory species. Also called the conservation development See limited develop-
Convention on the Conservation of Migratory Species ment. [11]
of Wild Animals. [6] conservation easements (CEs) Method of protect-
bushmeat Meat from any wild animal. [3] ing land in which landowners give up the right to
develop or build on their property, often in exchange
C for financial or tax benefit. [9]
carnivores An animal species that consumes other conservation leasing Providing payments to pri-
animals to survive. Also called a secondary consumer vate landowners who actively manage their land for
or predator. Compare with primary consumer. [2] biodiversity protection. [11]
carrying capacity The number of individuals or consumptive use value Value assigned to goods
biomass of a species that an ecosystem can support. [2] that are collected and consumed locally. [3]
census A count of the number of individuals in a Convention Concerning the Protection of
population. [6] the World Cultural and Natural Heritage
co-management Local people working as partners See World Heritage Convention. [11]
with government agencies and conservation organiza- Convention on Biological Diversity (CBD)
tions in protected areas. [9] A treaty that obligates countries to protect the biodi-
commodity values See direct use values. [3] versity within their borders, and gives them the right
common-property resources Natural resources to receive economic benefits from the use of that biodi-
that are not controlled by individuals but collec- versity. [11]
tively owned by society. Also known as open-access Convention on International Trade in Endan-
resources or common-pool resources. [3] gered Species (CITES) The international treaty that
community conserved areas Protected area man- establishes lists (known as Appendices) of species for
aged and sometimes established by local people. [9] which international trade is to be prohibited, regu-
lated, or monitored. [6]
community-based conservation (CBC) Protec-
tion of natural areas and resources that is controlled, Convention on the Conservation of Migratory
owned, and/or managed by the local people; an alter- Species of Wild Animals (CMS) See Bonn Conven-
native to government-based conservation. [9] tion. [6]
compensatory mitigation When a new site is cost–benefit analysis Comprehensive analysis that
created or rehabilitated in compensation for a site compares values gained against the costs of a project
damaged or destroyed elsewhere. Also known as or resource use. [3]
biodiversity offset. [10]
Glossary 417

cost–effectiveness analysis An alternative to cost– ecological resilience A natural ability to recover


benefit analysis that compares the impact (financial after disturbance. [10]
and otherwise) and costs of alternative means of ecological restoration Altering a site to reestablish
accomplishing an objective, such as the protection of a an indigenous ecosystem. [10]
species. [3] ecological succession A predictable, gradual and
cross-fostering Conservation strategy in which progressive change in species over time [8]
individuals from a common species raise the offspring ecologically extinct A species that has been so
of a rare, related species. [7] reduced in numbers that it no longer has a significant
cultural eutrophication Algal blooms and as- ecological impact on the biological community. See
sociated impacts caused by excess mineral nutrients functionally extinct. [5]
released into the water from human activity. [10] economic development Economic activity focused
D on improvements in efficiency and organization but
debt-for-nature swaps Agreements in which a de- not necessarily on increases in resource consumption.
veloping country agrees to fund additional conserva- [11]
tion activities in exchange for a conservation organiza- economic growth Economic activity characterized
tion canceling some of its discounted debt. [11] by increases in the amount of resources used and in
decomposers A species that feeds or grows on dead the amount of goods and services produced. [11]
plant and animal material. Also called a detritivore. [2] economics The study of factors affecting the pro-
deep ecology Philosophy emphasizing biodiversity duction, distribution, and consumption of goods and
protection, personal lifestyle changes, and working services. [3]
towards political change. [3] ecosystem diversity The variety of ecosystems
degazettement Government actions taken to present in a place or geographic area. [2]
remove the legal status of protected areas. [8] ecosystem engineers Species that modify the
demographic stochasticity Random variation in physical structure of an ecosystem. [2]
birth, death, and reproductive rates in small popula- ecosystem management Large-scale management
tions, sometimes causing further decline in population that often involves multiple stakeholders, the primary
size. Also called demographic variation. [5] goal of which is the preservation of ecosystem compo-
demographic studies Studies in which individuals nents and processes. [9]
and populations are monitored over time to determine ecosystem management Large-scale management
rates of growth, reproduction, and survival. [6] that often involves multiple stakeholders, the primary
demographic variation See demographic stochas- goal of which is the preservation of ecosystem compo-
ticity. [5] nents and processes. [1]
detritivores See decomposer. [2] ecosystem services Range of benefits provided to
people from ecosystems, including flood control, clean
direct use values Value assigned to products, such
water, and reduction of pollution.
as timber and animals, that are harvested and directly
used by the people who harvest them. Also known as ecosystem A biological community together with
commodity value or private goods. [3] its associated physical and chemical environment. [2]
ecotourism Tourism, especially in developing
E countries, focused on viewing unusual and/or espe-
Earth Summit An international conference held in cially charismatic biological communities and species
1992 in Rio de Janeiro that resulted in new environ- that are unique to a country or region. [3]
mental agreements. Also known as the Rio Summit. edge effects Altered environmental and biological
[11] conditions at the edges of a fragmented habitat. [4]
ecocolonialism Practice of governments and con- effective population size (Ne) The number of
servation organizations disregarding the land rights breeding individuals in a population. [5]
and traditions of local people in order to establish new
conservation areas. [9] embryo transfer The surgical implantation of
embryos into a surrogate mother; used to increase the
ecological economics Discipline that includes number of individuals of a rare species, with a com-
valuations of biodiversity in economic analyses. [3] mon species used as the surrogate mother. [7]
ecological footprint The influence a group of Endangered Species Act (ESA) An important
people has on both the surrounding environment US law passed to protect endangered species and the
and locations across the globe as measured by global ecosystems in which they live. [6]
hectares per person. [4]
418 Glossary

endemic species Species found in one place and extinction cascade A series of linked extinctions
nowhere else (e.g., the many lemur species found only whereby the extinction of one species leads to the
on the island of Madagascar). [2] extinction of one or more other species. [2]
endemic Occurring in a place naturally, without the extinction vortex Tendency of small populations to
influence of people (e.g., gray wolves are endemic to decline toward extinction. [5]
Canada). [5] extirpated Local extinction of a population, even
environmental economics Discipline that exam- though the species may still exist elsewhere. [5]
ines the economic impacts of environmental policies extractive reserve Protected area in which
and decisions. [3] sustainable extraction of certain natural products is
environmental ethics Discipline of philosophy that allowed. [9]
articulates the intrinsic value of the natural world and
people’s responsibility to protect the environment. [3] F
flagship species A species that captures pub-
environmental impact assessments Evaluation
lic attention, aids in conservation efforts, such as
of a project that considers its possible present and
establishing a protected area, and may be crucial to
future impacts on the environment. [3]
ecotourism. [6]
environmental justice Movement that seeks to
focal species A species that provides a reason for
empower and assist poor and politically weak people
establishing a protected area. [6]
in protecting their own environments; their well-being
and the protection of biological diversity are enhanced food chains Specific feeding relationships between
in the process. [3] species at different trophic levels. [2]
environmental stochasticity Random variation in food web A network of feeding relationships
the biological and physical environment. Can increase among species. [2]
the risk of extinction in small populations. [5] founder effect Reduced genetic variability that
environmentalism A widespread movement, occurs when a new population is established (“found-
characterized by political activism, with the goal of ed”) by a small number of individuals. [5]
protecting the natural environment. [1] four Rs Guidelines used by conservation biologists
eutrophication Process of degradation in aquatic when designing nature reserves: representation, resil-
environments caused by nitrogen and phosphorus iency, redundancy, and reality. [8]
pollution and characterized by algal blooms and oxy- frontier forest Intact blocks of undisturbed forest
gen depletion. [4] large enough to support all aspects of biodiversity. [4]
evolutionary definition of species A group of in- functional diversity The diversity of organisms
dividuals that share unique similarities of their DNA categorized by their ecological roles or traits rather
and hence their evolutionary past. [2] than their taxonomy. [2]
ex situ conservation Preservation of species under functionally extinct The state in which a species
artificial conditions, such as in zoos, aquariums, and persists at such reduced numbers that its effects on
botanical gardens. [7] the other species in its community are negligible. See
existence value The benefit people receive from ecologically extinct. [5]
knowing that a habitat or species exists and quanti- G
fied as the amount that people are willing to pay to gamma diversity The number of species in a large
prevent species from being harmed or going extinct, geographic area. [2]
habitats from being destroyed, and genetic variation
from being lost. [3] gap analysis Comparing the distribution of endan-
gered species and biological communities with exist-
extant Presently alive; not extinct. [5] ing and proposed protected areas to determine gaps in
externalities Hidden costs or benefits that result protection. [8]
from an economic activity to individuals or a society gene pool The total array of genes and alleles in a
not directly involved in that activity. [3] population. [2]
extinct in the wild A species no longer found in genes Units (DNA sequences) on a chromosome
the wild, but individuals may remain alive in zoos, that code for specific proteins. Also called loci. [2]
botanical gardens, or other artificial environments. [5]
genetic diversity The range of genetic variation
extinct The condition in which no members of a found within a species. [2]
species are currently living. [5]
genetic drift Loss of genetic variation and change
in allele frequencies that occur by chance in small
populations. [5]
Glossary 419

genetically modified organisms (GMOs) Organ- herbivores A species that eats plants or other
isms whose genetic code has been altered by scientists photosynthetic organisms. Also called a primary
using recombinant DNA technology. [4] consumer. [2]
genome resource banking (GRB) Collecting DNA, herbivory Predation on plants. [2]
eggs, sperm, embryos, and other tissues from species heterozygous Condition of an individual having
that can be used in breeding programs and scientific two different allele forms of the same gene. [2]
research. [7] homozygous Condition of an individual having
genotype Particular combination of alleles that an two identical allele forms of the same gene. [2]
individual possesses. [2] hotspots Regions with numerous species, many
geographic information systems (GIS) Computer of which are endemic, that are also under immediate
analyses that integrate and display spatial data; relat- threat from human activity. [6]
ing in particular to the natural environment, ecosys- hybridize Interbreeding between different spe-
tems, species, protected areas, and human activities. [8] cies. [2]
Global Environment Facility (GEF) A large inter- hybrids Intermediate offspring resulting from mat-
national program involved in funding conservation ing between individuals of two different species. [2]
activities in developing countries. [11]
globalization The increasing interconnectedness of I
the world’s economy. [4] in situ conservation Preservation of natural com-
globally extinct No individuals are presently alive munities and populations of endangered species in
anywhere. [5] the wild. [7]
greenhouse effect Warming of the Earth caused by inbreeding depression Lowered reproduction
carbon dioxide and other “greenhouse gases” in the or production of weak offspring following mating
atmosphere that allow the sun’s radiation to penetrate among close relatives or self-fertilization. [5]
and warm the Earth but prevent the heat generated by indicator species Species used in a conservation
sunlight from re-radiating. Heat is thus trapped near plan to identify and often protect a biological commu-
the surface, raising the planet’s temperature. [4] nity or set of ecosystem processes. [6]
greenhouse gases Gases in the atmosphere, indirect use values Values provided by biodiver-
primarily carbon dioxide, that are transparent to sun- sity that do not involve harvesting or destroying the
light but that trap heat near the Earth’s surface. [4] resource (such as water quality, soil protection, recre-
guild A group of species at the same trophic level ation, and education). Also known as public goods. [3]
that use approximately the same environmental integrated conservation development projects
resources. [2] (ICDPs) Conservation projects that also provide for
the economic needs and welfare of local people. [9]
H
integrated pest management An approach to
habitat conservation plans (HCPs) Regional controlling undesirable plants or animals that has the
plans that allow development in designated areas goal of minimizing harm to the ecosystem and people,
while protecting biodiversity in other areas. [6] while being cost-effective. [4]
habitat corridors See conservation corridor. [8] International Union for Conservation of
habitat fragmentation The process whereby a Nature (IUCN) See IUCN. [4]
large, continuous area of habitat is both reduced in intrinsic value Value of a species and other aspects
area and divided into two or more fragments. [4] of biodiversity for their own sake, unrelated to human
habitat islands Intact habitat surrounded by an needs. [3]
unprotected matrix of inhospitable terrain. [8] introduction program Moving individuals to areas
habitat The location or type of environment in outside their historical range in order to create a new
which a specific animal or plant species lives. [2] population of an endangered species. [7]
hard release In the establishment of a new popula- invasive An introduced species that increases in
tion, when individuals from an outside source are abundance at the expense of native species. [4]
released in a new location without assistance. Compare island biogeography model Formula for the
with soft release. [7] relationship between island size and the number of
healthy ecosystem Ecosystem in which processes species living on the island; the model can be used to
are functioning normally, whether or not there are hu- predict the impact of habitat destruction on species
man influences. [2] extinctions, viewing remaining habitat as an “island”
in the “sea” of a degraded ecosystem. [5]
420 Glossary

IUCN International Union for the Conservation of market failure Misallocation of resources in which
Nature is a major international conservation organiza- certain individuals or businesses benefit from using a
tion; previously known as The World Conservation common resource, such as water, the atmosphere, or a
Union. [4] forest, but other individuals, businesses or the society
at large bears the cost. [3]
K
metapopulation Shifting mosaic of populations of
keystone resources Any resource in an ecosystem
the same species linked by some degree of migration;
that is crucial to the survival of many species; for
a “population of populations.” [6]
example, a watering hole. [2]
minimum dynamic area (MDA) Area needed for
keystone species A species that has a dispropor-
a population to have a high probability of surviving
tionate impact (relative to its numbers or biomass) on
into the future. [6]
the organization of a biological community. Loss of a
keystone species may have far-reaching consequences minimum viable population (MVP) Number of
for the community. [2] individuals necessary to ensure a high probability that
a population will survive a certain number of years
L into the future. [6]
land ethic Aldo Leopold’s philosophy advocating mitigation Process by which a new population or
human use of natural resources that is compatible habitat is created to compensate for a habitat damaged
with or even enhances ecosystem health. [1] or destroyed elsewhere. [7]
land sharing Land use which combines resource morphological definition of species A group of
use and conservation. [9] individuals, recognized as a species, that is morpho-
land sparing Land which is protected when other logically, physiologically, or biochemically distinct
lands are used more intensively. [9] from other groups. Compare with biological definition
land trusts Conservation organizations that protect of species. [2]
and manage land. [11] morphospecies Individuals that are probably a
landraces A variety of crop that has unique genetic distinct species based on their appearance but that do
characteristics; local species that have been adapted by not currently have a scientific name. [2]
humans over time. [7] movement corridors See conservation corridors.
landscape ecology Discipline that investigates [8]
patterns of habitat types and their influence on species multiple-use habitat An area managed to provide
distribution and ecosystem processes. [8] a variety of goods and services. [9]
legal title The right of ownership of land, rec- mutations Changes that occur in genes and chro-
ognized by a government and/or judicial system; mosomes, sometimes resulting in new allele forms
traditional people often struggle to achieve this and genetic variation. [2]
recognition. [9] mutualism When two species benefit each other by
limited development Compromise involving a their relationship. [2]
landowner, a property developer, and a conservation
organization that combines some development with N
protection of the remaining land. [11] national environmental fund (NEF) A trust fund
or foundation that uses its annual income to support
limiting resource Any requirement for existence
conservation activities. [11]
whose presence or absence limits a population’s size. In
the desert, for example, water is a limiting resource. [2] natural history The ecology and distinctive charac-
teristics of a species. [6]
Living Planet Index A measure of the conservation
status of species, based on the IUCN categories. [6] non-use values Values of something that is not
presently used; for example, existence value. [3]
locally extinct A species that no longer exists in a
place where it used to occur, but still exists else- nonconsumptive use value Value assigned to
where. [5] benefits provided by some aspect of biodiversity that
does not involve harvesting or destroying the resource
loci (singular, locus) See genes. [2]
(such as water quality, soil protection, recreation, and
M education). [3]
management plans A statement of how to protect nonpoint source pollution Pollution coming from
biodiversity in an area, along with methods for imple- a general area rather than a specific site. [10]
mentation. [8] normative discipline A discipline that embraces
ethical commitment rather than ethical neutrality. [1]
Glossary 421

novel ecosystems Ecosystems in which there is a precautionary principle Principle stating that it
mixture of native and nonnative species coexisting in may be better to avoid taking a particular action due
a community unlike the original or reference site. [10] to the possibility of causing unexpected harm. [3]
O predation Act of killing and consuming another
organism for food. [2]
omnivores Species that eat both plants and ani-
mals. [2] predator release hypothesis An hypothesis
which attributes the success of invasive species to the
open-access resources Natural resources that
absence of specialized natural predators and parasites
are not controlled by individuals but are collectively
in their new range. [4]
owned by society. [3]
predators See carnivores. [2]
option value Value of biodiversity in providing
possible future benefits for human society (such as preservationist ethic A belief in the need to pre-
new medicines). [3] serve wilderness areas for their intrinsic value. [1]
prey An animal that is eaten as food by another
P species. [2]
parasites Organisms that live on or in another primary consumers See herbivores. [2]
organism (host), receiving nutritive benefit while de-
creasing the fitness of the host, which remains alive. [2] primary producers Organisms such as green
plants, algae, and seaweeds that obtain their energy
Paris Accord An agreement made in Paris in 2015 directly from the sun via photosynthesis. Also known
by 195 nations to lower greenhouse gas emissions as autotrophs. [2]
with the goal of preventing atmospheric temperatures
from increasing more than 2°C. [11] private goods See direct use values. [3]
passive restoration Letting an ecosystem recover productive use value Value assigned to products
on its own. [10] that are sold in markets. [3]
pathogens Disease-causing organisms. [2] protected area A habitat managed primarily or in
large part for biodiversity. [8]
payments for ecosystem services (PES) Direct
payments to individual landowners and local com- public goods Nonconsumptive benefits that belong
munities that protect species or critical ecosystem to society in general, without private ownership. Also
characteristics. [9] known as indirect use values. [3]
perverse subsidies Government payments or other R
financial incentives to industries that result in environ- Ramsar Convention on Wetlands A treaty that
mentally destructive activities. [3] promotes the protection of wetlands of international
phenotype The morphological, physiological, importance. [11]
anatomical, and biochemical characteristics of an indi- rapid biodiversity assessments Species invento-
vidual that result from the expression of its genotype ries and vegetation maps made by teams of biologists
in a particular environment. [2] when urgent decisions must be made on where to
polymorphic genes Within a population, genes establish new protected areas. Also known as rapid
that have more than one form or allele. [2] assessment plans (RAPs). [6]
population biology Study of the ecology and ge- recombination Mixing of the genes on the two cop-
netics of populations, often with a focus on population ies of a chromosome that occurs during meiosis (i.e.,
numbers. [6] in the formation of egg and sperm, which contain only
population bottleneck A radical reduction in one copy of each chromosome). Recombination is an
population size (e.g., following an outbreak of infec- important source of genetic variation. [2]
tious disease), sometimes leading to the loss of genetic reconciliation ecology The science of developing
variation. [5] urban places in which people and biodiversity can
population viability analysis (PVA) Demographic coexist. [9]
analysis that predicts the probability of a population reconciliation ecology The science of developing
persisting in an environment for a certain period urban places in which people and biodiversity can
of time; sometimes linked to various management coexist. [10]
scenarios. [6] recovery criteria Predetermined thresholds (such
population A geographically defined group of in- as numbers of individuals alive in the wild) that signal
dividuals of the same species that mate and otherwise that an endangered species can be removed from pro-
interact with one another. Compare with metapopula- tection under the Endangered Species Act. [6]
tion. [2]
422 Glossary

Red Data Books Compilations of lists (“Red Lists”) Shannon diversity index A species diversity index
of endangered species prepared by the IUCN and that takes into account the numbers of different spe-
other conservation organizations. [6] cies and their relative abundance. [2]
Red List criteria Quantitative measures of threats shifting cultivation Farming method in which
to species based on the probability of extinction. [6] farmers cut down trees, burn them, plant crops for a
Red List Index Measure of the conservation status few years, and then abandon the site when soil fertility
of species based on the IUCN categories. [6] declines. Also called “slash-and-burn” agriculture. [4]
Red Lists Lists of endangered species prepared by sink populations Populations that receive an influx
the IUCN. [6] of new individuals from a source population. [6]
Reducing Emissions from Deforestation and For- sixth extinction episode The present mass extinc-
est Degradation (REDD) Program using financial tion event which is just beginning. [5]
incentives to reduce the emissions of greenhouse gases SLOSS debate Controversy concerning the relative
from deforestation. [11] advantages of a single large or several small conserva-
reference sites Control site that provides goals for tion areas. “SLOSS” stands for single large or several
restoration in terms of species composition, commu- small. [8]
nity structure, and ecosystem processes. [10] soft release In the establishment of a new popula-
reinforcement program Releasing new individuals tion, when individuals are given assistance during
into an existing population to increase population size or after the release to increase the chance of success.
and genetic variability. [7] Compare with hard release. [7]
reintroduction program The release of captive source populations Established populations from
bred or wild-collected individuals at a site within their which individuals disperse to new locations. [6]
historical range where the species does not presently species diversity The entire range of species found
occur. [7] in a particular place. [2]
replacement cost approach How much people species richness The number of species found in a
would have to pay for an equivalent product if what community. [2]
they normally use is unavailable. [3] species–area relationship The number of species
representative site Protected area that includes found in an area increases with the size of the area;
species and ecosystem properties characteristic of a i.e., more species are found on large islands than on
larger area. [6] small islands. (5)
resilience The ability of an ecosystem to return to stable ecosystems Ecosystems that are able to
its original state following disturbance. [2] remain in roughly the same compositional state de-
resistance The ability of an ecosystem to remain in spite human intervention or stochastic events such as
the same state even with ongoing disturbance. [2] unseasonable weather. [2]
resource conservation ethic Natural resources stochasticity Random variation; variation happen-
should be used for the greatest good of the largest ing by chance. [5]
number of people for the longest time. [1] survey Repeatable sampling method to estimate
restoration ecology The scientific study of restored population size or density, or some other aspect of
populations, communities, and ecosystems. [10] biodiversity. [6]
rewilding Returning species, in particular large sustainable development Economic development
mammals to landscape, to approximate their natural that meets present and future human needs without
condition prior to human impact. [10] damaging the environment and biodiversity. [1]
Rio Summit See Earth Summit. [11] symbiotic A mutualistic relationship in which nei-
ther of the two species involved can survive without
S the other. [2]
secondary consumers See carnivores. [2]
T
secondary invasion When the removal of an inva-
sive species is followed by an invasion by a different taxonomists Scientists involved in the identifica-
species. [10] tion and classification of species. [2]
seed banks Collections of stored seeds, collected tertiary consumers The fourth trophic level, in
from wild and cultivated plants; used in conservation which predators eat other predators. [2]
and agricultural programs. [7] threatened Species that fall into the endangered
or vulnerable to extinction categories in the IUCN
system. Under the US Endangered Species Act, refers
Glossary 423

to species at risk of extinction, but at a lower risk than use values The direct and indirect values provided
endangered species. [6] by some aspect of biodiversity. [3]
tragedy of the commons The unregulated use of W
a public resource that results in its degradation. [3]
World Bank International bank established to
transects Lines often designated with measuring support economic development in developing
tape or permanent markers, along which biological countries. [11]
data is collected. [6]
World Heritage Convention (WHC) A treaty that
trophic cascade Major changes in vegetation and protects cultural and natural areas of international
biodiversity resulting from the loss of a keystone spe- significance. [11]
cies. [2]
World Heritage site A cultural or natural area
trophic levels Levels of biological communities officially recognized as having international signifi-
representing ways in which energy is captured and cance. [8]
moved through the ecosystem by the various types
World Summit on Sustainable Development
of species. See primary producer; herbivore; predator;
Held in Johannesburg, South Africa, in 2002, this gath-
detritivore. [2]
ering emphasized achieving the social and economic
U goals of sustainability. [11]
umbrella species Protecting an umbrella species Z
results in the protection of other species. [6]
zoning A method of managing protected areas that
UN Conference on Sustainable Development allows or prohibits certain activities in designated
Held in 2012, this conference linked biodiversity places. [8]
conservation to sustainable development and control-
ling climate change, and emphasized the need for
market-based solutions. Also unofficially called the
Rio+20. [11]
Chapter Opening
Photograph Credits
Chapter 1 Photograph by Hector R. Chenge, courtesy of Gladys Porter Zoo.
Chapter 2 © pnup65/Getty Images.
Chapter 3 © Michael Myers.
Chapter 4 © trekkerimages/Alamy Stock Photo.
Chapter 5 © Torsten Blackwood/AFP/Getty Images.
Chapter 6 Photograph by Scott Dressel-Martin, Denver Botanic Gardens.
Chapter 7 © blickwinkel/Alamy Stock Photo.
Chapter 8 Photograph by John Fielder.
Chapter 9 © Martin Harvey/Getty Images.
Chapter 10 © Aman Rochman/AFP/Getty Images.
Chapter 11 © Paolo Patrizi/Alamy Stock Photo.
Chapter 12 Photograph by Scott Dressel-Martin, Denver Botanic Gardens.
Bibliography
Numbers in parentheses indicate the chapter(s) in which the reference is cited.

Abesamis, R. A. and G. R. Russ. 2005. Density-dependent Alroy, J. 2015. Limits to captive breeding of mammals in
spillover from a marine reserve: Long-term evi- zoos. Conservation Biology 29(3): 926–931. (7)
dence. Ecological Applications 15: 1798–1812. (8) Alter, S. E., E. Rynes, and S. R. Palumbi. 2007. DNA evi-
Abson, D. J. and M. Termansen. 2011. Valuing eco-system dence for historic population size and past ecosystem
services in terms of ecological risks and returns. impacts of gray whales. Proceedings of the National
Conservation Biology 25: 250–258. (3) Academy of Sciences USA 104: 15162–15167. (4)
Acuña, V. and A. Ruhí. 2016. Temporary streams: current Altman, I., A. M. H. Blakeslee, G. C. Osio, C. B. Rillahan,
management challenges and promising solutions. et al. 2011. A practical approach to implementation
Restoration Ecology. In press. (8) of ecosystem-based management: A case study us-
Adams, J. S. 2006. The Future of the Wild: Radical Conserva- ing the Gulf of Maine marine ecosystem. Frontiers in
tion for a Crowded World. Beacon Press, Boston. (10) Ecology and the Environment 9: 183–189. (9)
Adum, G. B., M. P. Eichhorn, W. Oduro, C. Ofori- Alva-Basurto, J. C. and J. Arias-Gonzalez. 2014. Model-
Boateng, et al. 2013. Two-stage recovery of amphib- ling the effects of climate change on a Caribbean
ian assemblages following selective logging of coral reef food web. Ecological Modelling 289: 1–14.
tropical forests. Conservation Biology 27: 354–363. (9) (2)
Aguirre-Acosta, N., E. Kowaljow, and R. Aguilar. 2014. Anadón, J. D., A. Gimenez, R. Ballestar, and I. Pérez.
Reproductive performance of the invasive tree 2009. Evaluation of local ecological knowledge as
Ligustrum lucidum in a subtropical dry forest: Does a method for collecting extensive data on animal
habitat fragmentation boost or limit invasion?. abundance. Conservation Biology 23: 617–625. (8)
Biological Invasions 16(7): 1397–1410. (4) Andam, K. S., P. J. Ferraro, K. R. E. Sims, A. Healy, and
Albert, A., K. McKonkey, T. Savini, and M. C. Huynen. M. B. Holland. 2010. Protected areas reduced pov-
2014. The value of disturbance-tolerant cercopi- erty in Costa Rica and Thailand. Proceedings of the
thecine monkeys as seed dispersers in degraded National Academy of Sciences USA 107: 9996–10001.
habitats. Biological Conservation 170: 300–310. (2) (8)
Alexander, S. (ed.). 2009. Voluntary Simplicity: The Poetic Anderson, M. G. and C. E. Ferree. 2010. Conserving the
Alternative to Consumer Culture. Stead and Daugh- stage: climate change and the geophysical under-
ters, Whanganui, New Zealand. (3) pinnings of species diversity. PLoS ONE 5: e11554.
(8)
Allee, W. C. 1931. Animal aggregations: a study in
general sociology. The University of Chicago Press, Andersson, E., S. Barthel, and K. Ahrne. 2007. Measuring
Chicago. (5) social-ecological dynamics behind the generation
of ecosystem services. Ecological Applications. 17:
Allen, C., R. S. Lutz, and S. Demarais. 1995. Red import-
1267–1287. (8)
ed fire ant impacts on northern bobwhite popula-
tions. Ecological Applications 5: 632–638. (4) Andrello, M., M. N. Jacobi, S. Manel, W. Thuiller, and D.
Mouillot. 2015. Extending networks of protected ar-
Allen, L. S. 2006. Collaboration in the borderlands: the
eas to optimize connectivity and population growth
Malpai Borderlands Group. Rangelands 28(3): 17–21.
rate. Ecography 38(3): 273–282. (8)
(9)
Andrew-Essien, E. and F. Bisong. 2009. Conflicts, conser-
Allen, W. H. 1988. Biocultural restoration of a tropical
vation and natural resource use in protected area
forest: architects of Costa Rica’s emerging Guana-
systems: An analysis of recurrent issues. European
caste National Park plan to make it an integral part
Journal of Scientific Research 25: 118–129. (9)
of local culture. BioScience 38: 156–161. (10)
Angelsen, A., P. Jagger, R. Babigumira, B. Belcher, et al.
Allendorf, F. W. and G. Luikart. 2007. Conservation and
2014. Environmental income and rural livelihoods:
the Genetics of Populations. Blackwell Publishing,
A global-comparative analysis. World Development
Oxford, UK. (5)
64: S12–S28. (3)
Allendorf, T. D. and K. Allendorf. 2012. What every con-
Área de Conservación Guanacaste (ACG). www.acguana-
servation biologist should know about human popu-
caste.ac.cr. (10)
lation. Conservation Biology 26(6): 1523–1739. (4)
426 Bibliography

Arima, E. Y., R. T. Walker, S. Perz, and C. Souza, Jr. Balmford, A., A. Bruner, P. Cooper, R. Costanza, et al.
2015. Explaining the fragmentation in the Brazilian 2002. Economic reasons for conserving wild nature.
Amazonian forest. Journal of Land Use Science. doi: Science 297: 950–953. (3)
10.1080/1747423X.2015.1027797. (5) Ban, N. C., M. Mills, J. Tam, C. C. Hicks, et al. 2013. A
Arizmendi-Mejía, R., C. Linares, J. Garrabou, A. Antunes, social-ecological approach to conservation plan-
et al. 2015. Combining Genetic and Demographic ning: Embedding social considerations. Frontiers in
Data for the Conservation of a Mediterranean Ecology and the Environment 11: 194–202. (9)
Marine Habitat-Forming Species. PLoS ONE, 10(3), Banerjee, S., S. Secchi, J. Fargione, S. Polasky, et al.
e0119585. (6) 2013. How to sell ecosystem services: a guide for
Armstrong, D. P. and P. J. Seddon. 2008. Directions in designing new markets. Frontiers in Ecology and the
reintroduction biology. Trends in Ecology and Evolu- Environment 11: 297–304. (9)
tion 23: 20–25. (7) Banga, S. S. and M. S. Kang. 2014. Developing climate-re-
Armsworth, P. R., G. C. Daily, P. Kareiva, and J. N. silient crops. Journal of Crop Improvement 28: 57–87. (7)
Sanchirico. 2006. Land market feedbacks can under- Barazani, O., A. Perevolotsky, and R. Hadas. 2008. A
mine biodiversity conservation. Proceedings of the problem of the rich: prioritizing local plant genetic
National Academy of Sciences USA 103: 5403–5408. (8) resources for ex situ conservation in Israel. Biological
Arthington, A. H., J. M. Bernardo, and M. Ilhéu. 2014. Conservation 141: 596–600. (9)
Temporary rivers: Linking ecohydrology, ecological Barber-Meyer, S. M. 2010. Dealing with the clandestine
quality and reconciliation ecology. River research and nature of wildlife-trade market surveys. Conserva-
Applications 30(10): 1209–1215. (10) tion Biology 24: 918–923. (3)
Association of Zoos and Aquariums. 2009. www.aza.org (7) Barbier, E. B., J. C. Burgess, J. T. Bishop, and
Athreya, V., M. Odden, J. D. Linnell, J. Krishnaswamy, B. A. Aylward. 1994. The economics of the tropical
and U. Karanth. 2013. Big cats in our backyards: timber trade. Earthscan Publications, London. (3)
Persistence of large carnivores in a human domi- Barcia, L. 2015. Addis Ababa: Financing the future or fi-
nated landscape in India. PLoS ONE 8(3): e57872. (9) nancing failure? devex. www.devex.com/news/addis-aba-
Audino, L. D., J. Louzada, and L. Comita. 2014. Dung ba-financing-the-future-or-financing-failure-86561. (11)
beetles as indicators of tropical forest restoration: is Barnosky, A. D., N. Matzke, S. Tomiya, G. O. U. Wogan,
it possible to recover species and functional diver- et al. 2011. Has Earth’s sixth mass extinction already
sity? Biological Conservation 169: 248–257. (10) arrived? Nature 471: 51–57. (1)
Azam, F. and A. Z. Worden. 2004. Oceanography: Mi- Baskin, Y. 1997. The Work of Nature: How the Diversity of
crobes, molecules, and marine ecosystems. Science Life Sustains Us. Island Press, Washington, DC. (3)
303: 1622–1624. (2)
Basset, Y., H. Barrios, S. Segar, R. B. Srygley, et al. 2015.
The Butterflies of Barro Colorado Island, Panama:
Bagla, P. 2010. Hardy cotton-munching pests are latest Local Extinction since the 1930s. PLoS ONE 10(8):
blow to GM crops. Science 327: 1439. (4) e0136623. (5)
Baillie, J. E., B. Collen, R. Amin, H. R. Akcakaya, H. R., Bateman, H. L., D. M. Meritt, E. P. Glenn, P. L. Nagler.
et al. 2008. Toward monitoring global biodiversity. 2014. Indirect effects of biocontrol of an invasive
Conservation Letters 1(1): 18–26. (6) riparian plant (Tamarix) alters habitat and reduces
Baker, C. M., M. Bode, and M. A. McCarthy. 2016. Models herpetofauna abundance. Biological Invasions 17:
that predict ecosystem impacts of reintroductions 87–97. (10)
should consider uncertainty and distinguish be- Bateman, I. J., A. R. Harwood, G. M. Mace, R. T. Wat-
tween direct and indirect effects. Biological Conserva- son, et al. 2013. Bringing ecosystem services into
tion. doi: 10.1016/j.biocon.2016.01.023. (7) economic decision-making: land use in the United
Baker, J., E. J. Milner-Gulland, and N. Leader-Williams. Kingdom. Science 341: 45–50. (3, 4)
2012. Park gazettement and integrated conservation Baudron, F. and K. E. Giller. 2014. Agriculture and na-
and development as factors in community conflict ture: Trouble and strife? Biological Conservation 170:
at Bwindi Impenetrable Forest, Uganda. Conserva- 232–245. (9)
tion Biology 26: 160–170. (9) Bayraktarov, E., M. I. Saunders, S. Abdullah, M. Mills, et
Baker, J. D. and P. M. Thompson. 2007. Temporal and al. 2015. The cost and feasibility of marine coastal
spatial variation in age-specific survival rates of a restoration. Ecological Applications. (10)
long-lived mammal, the Hawaiian monk seal. Pro- Bean, D., T. Dudley, and K. Hultine. 2013. Bring on the
ceedings of the Royal Society B 274: 407–415. (6) beetles! The history and impact of tamarisk biologi-
Balmford, A. 1996. Extinction filters and current resil- cal control. In A. Sher and M. F. Quigley (eds.),
ience: the significance of past selection pressures for Tamarix: A Case Study of Ecological Change in the
conservation biology. TREE 11: 193–196. (5) American West,
Balmford, A., J. Beresford, J. Green, R. Naidoo, et al. 2009. pp. 377–403. Oxford University Press. (10)
A global perspective on trends in nature-based tour- Beane, J. C., S. P. Graham, T. J. Thorp, and L. T. Pusser.
ism. PLoS Biology 7: e1000144. (3) 2014. Natural history of the southern hognose snake
Bibliography 427

(Heterodon simus) in North Carolina, USA. Copeia 1: Kemp’s ridley sea turtle, Lepidochelys kempii. Herpe-
168–175. (6) tological Conservation and Biology 9: 563–577. (1)
Beans, C. M. and D. A. Roach. 2015. An invasive plant Beyer, H. L., E. H. Merrill, N. Varley, and M. S. Boyce.
alters pollinator-mediated phenotypic selection on 2007. Willow on Yellowstone’s northern range: Evi-
a native congener. American Journal of Botany 102: dence for a trophic cascade? Ecological Applications
50–57. (4) 17: 1563–1571. (2)
Bearzi, G. 2009. When swordfish conservation biologists Bhagwat, S. A., N. Dudley, and S. R. Harrop. 2011. Reli-
eat swordfish. Conservation Biology 23: 1–2. (3) gious following in biodiversity hotspots: Challenges
Beattie, A. and P. Ehrlich. 2010. The missing link in biodi- and opportunities for conservation and develop-
versity conservation. Science 328: 307–308. (3) ment. Conservation Letters 4: 234–240. (3)
Becker, C. G., C. R. Fonseca, C. F. B. Haddad, and P. I. Bhagwat, S. A., S. Nogué, and K. J. Willis, K. J. 2012. Resil-
Prado. 2010. Habitat split as a cause of local popula- ience of an ancient tropical forest landscape to 7500
tion declines of amphibians with aquatic larvae. years of environmental change. Biological Conserva-
Conservation Biology 24: 287–294. (4) tion 153: 108–117. (2)
Becker, M., R. McRobb, F. Watson, E. Droge, et al. 2013. Bhattacharya, A., J. Oppenheim, and N. Stern. 2015.
Evaluating wire-snare poaching trends and the im- Driving sustainable development through better
pacts of by-catch on elephants and large carnivores. infrastructure: Key elements of a transformation
Biological Conservation 158: 26–36. (8) program. Global Economy and Development Working
Paper 91. (11)
Beebee, T. J. C. 2013. Effects of road mortality and mitiga-
tion measures on amphibian populations. Conserva- Bhatti, S., S. Carrizosa, P. McGuire, and T. Young (eds.).
tion Biology 27: 657–668. doi: 10.1111/cobi.12063. (4) 2009. Contracting for ABS: The Legal and Scientific
Implications of Bioprospecting Contracts. IUCN, Gland,
Beier, P., W. Spencer, R. F. Baldwin, and B. H. McRae. 2011.
Switzerland. (3)
Towards best practices for developing regional con-
nectivity maps. Conservation Biology 25: 879–892. (8) Bickford, D., M. R. C. Posa, L. Qie, A. Campos-Arceiz,
et al. 2012. Science communication for biodiversity
Beissinger, S. R. 2015. Endangered species recovery cri-
conservation. Biological Conservation 151: 74–76. (12)
teria: reconciling conflicting views. BioScience 65(2):
121–122. (6) Bisht, I. S., P. S. Mehta, and D. C. Bhandari. 2007. Tradi-
tional crop diversity and its conservation on-farm
Beissinger, S. R., E. Nicholson, and H. P. Possingham.
for sustainable agricultural production in Kumaon
2009. Application of population viability analysis to
Himalaya of Uttaranchal state: A case study. Genetic
landscape conservation planning. In J. J. Millspaugh
Resources and Crop Evolution 54: 345–357. (9)
and F. R. Thompson, III (eds.), Models for Planning
Wildlife Conservation in Large Landscapes, pp. 33–50. Blackwell, M. 2011. The Fungi: 1, 2, 3… 5.1 million spe-
Academic Press, San Diego, CA. (6) cies? American Journal of Botany 98(3): 426–438. (2)
Bell, C. D., J. M. Blumenthal, A. C. Broderick, and B. J. Blanco, G., J. A. Lemus, and M. Garía-Montijano. 2011.
Godley. 2010. Investigating potential for depensa- When conservation management becomes contrain-
tion in marine turtles: How low can you go? Conser- dicated: Impact of food supplementation on health
vation Biology 24: 226–235. (5) of endangered wildlife. Ecological Applications 21:
2469–2477. (7)
Bennett, A. F. 1999. Linkages in the Landscape: The Role of
Corridors and Connectivity in Wildlife Conservation. Blickley, J. L., K. Deiner, K. Garbach, I. Lacher, et al. 2013.
IUCN, Gland, Switzerland. (8) Graduate student’s guide to necessary skills for
nonacademic conservation careers. Conservation
Berger, J. 1990. Persistence of different-sized populations:
Biology 27: 24–34. (12)
An empirical assessment of rapid extinctions in
bighorn sheep. Conservation Biology 4: 91–98. (6) Bobbink, R., K. Hicks, J. Galloway, T. Spranger, et al.
2010. Global assessment of nitrogen deposition
Berger, J. 1999. Intervention and persistence in small
effects on terrestrial plant diversity. Ecological Ap-
populations of bighorn sheep. Conservation Biology
plications 20: 30–59. (4)
13: 432–435. (6)
Bocci, A., S. Menapace, S. Alemanno, and S. Lovari. 2016.
Berger, J., B. Buuveibaatar, and C. Mishra. 2013. Global-
Conservation introduction of the threatened Apen-
ization of the cashmere market and the decline of
nine chamois Rupicapra pyrenaica ornata: post-release
large mammals in Central Asia. Conservation Biology,
dispersal differs between wild-caught and captive
27: 679–689. (4)
founders. Oryx 50(01): 128–133. (7)
Berger-Tal, O., D. T. Blumstein, S. Carroll, R. N. Fischer, et
Bode, M., W. Probert, W. R. Turner, K. A. Wilson, and O.
al. 2016. A systematic survey of the integration of be-
Venter. 2011. Conservation planning with multiple
havior into wildlife conservation and management.
organizations and objectives. Conservation Biology
Conservation Biology. doi: 10.1111/cobi.12654. (7)
25: 295–304. (8, 11)
Bevan, E., T. Wibbels, B. M. Z. Najera, M. A. C. Martinez,
Boersma, P. D. 2006. Landscape-level conservation for the
et al. 2014. In situ nest and hatchling survival at
sea. In M. J. Groom, G. K. Meffe, and C. R. Carroll
Rancho Nuevo, the primary nesting beach of the
(eds.), Principles of Conservation Biology, 3rd Ed., pp.
447–448. Sinauer Associates, Sunderland, MA. (6)
428 Bibliography

Boersma, P. D. and G. A. Rebstock. 2009. Foraging ogy: A Synthetic Approach to Ecological Research, pp.
distance affects reproductive success in Magel- 53–74. Cambridge University Press, Cambridge. (10)
lanic penguins. Marine Ecology Progress Series 375: Bradshaw, C. J. A. and B. W. Brook. 2014. Human popula-
263–275. (6) tion reduction is not a quick fix for environmental
Boitani, L., P. Ciucci, and E. Raganella-Pelliccioni. 2011. problems. Proceedings of the National Academy of
Ex-post compensation payments for wolf predation Sciences 111: 16610–16615. (1, 4)
on livestock in Italy: a tool for conservation?.Wildlife Bradshaw, C. J. A., B. M Fitzpatrick, C. C Steinberg, B. W.
Research 37(8): 722–730. (7) Brook, et al. 2008. Decline in whale shark size and
Bombaci, S. P., C. M. Farr, H. T. Gallo, A. M. Mangan, et abundance at Ningaloo Reef over the past decade:
al. 2016. Using Twitter to communicate conservation the world’s largest fish is getting smaller. Biological
science from a professional conference. Conservation Conservation 141: 1894–1190. (6)
Biology 30(1): 216–225. (12) Bradshaw, C. J. A., N. S. Sodhi, and B. W. Brook. 2009.
Bonsall, M. B., C. A. Dooley, A. Kasparson, T. Brereton, et Tropical turmoil: A biodiversity tragedy in progress.
al. 2014. Allee effects and the spatial dynamics of a Frontiers in Ecology and the Environment 7: 79–87. (4)
locally endangered butterfly, the high brown fritil- Bragina, E. V., V. C. Radeloff, M. Baumann, K. Wendland,
lary. Ecological Applications 24: 108–120. (5) et al. 2015. Effectiveness of protected areas in the
Borrini-Feyerabend, G., M. Pimbert, T. Farvar, A. Kothari, Western Caucasus before and after the transition to
and Y. Renard. 2004. Sharing Power: Learning by Do- post-socialism. Biological Conservation 184: 456–464. (8)
ing in Co-Management of Natural Resources through- Braithwaite, R. W. 2001. Role of Tourism. In S. A.
out the World. IIED and IUCN/CEESP/CMWG, Levin (ed.), Encyclopedia of Biodiversity, Vol. 5, pp.
Cenesta, Tehran. (9) 667–679. Academic Press, San Diego, CA. (3)
Botanic Gardens Conservation International (BGCI). Branton, M. and J. S. Richardson. 2011. Assessing the
2005. www.bgci.org (7) value of the umbrella-species concept for conser-
Botts, E. A., B. F. N. Erasmus, and G. J Alexander. 2013. vation planning with meta-analysis. Conservation
Small range size and narrow niche breadth predict Biology 25: 9–20. (6)
range contractions in South African frogs. Global Braschler, B. 2009. Successfully implementing a citi-
Ecology and Biogeography 22: 567–576. (5) zen-scientist approach to insect monitoring in a
Bouché, P., I. Douglas-Hamilton, G. Wittemyer, A. J. resource-poor country. BioScience 59: 103–104. (9)
Nianogo, et al. 2011. Will elephants soon disap- Braunisch, V., Coppes, J., Bachle, S., and Suchant, R. 2015.
pear from West African savannahs? PLoS ONE 6(6): Underpinning the precautionary principle with
e20619. (6) evidence: A spatial concept for guiding wind power
Bouwman, H., R. Bornman, C. Van Dyk, and I. Barn- development in endangered species’ habitats. Jour-
hoorn. 2015. First report of the concentrations and nal for Nature Conservation 24: 31–40. (3)
implications of DDT residues in chicken eggs from a Braunisch, V., S.-L. Huang, Y. Hao, S. T. Turvey, et al.
malaria-controlled area. Chemosphere 137: 174–177. (4) 2012. Conservation science relevant to action: a
Bowen, B. W., L. A. Rocha, R. J. Toonen, S. A. Karl, and research agenda identified and prioritized by practi-
the ToBo Laboratory. 2013. The origins of tropical tioners. Biological Conservation 153: 201–210. (8)
marine biodiversity. Trends in Ecology and Evolution Briggs, S. V. 2009. Priorities and paradigms: Directions in
28: 359–366. Patterns and processes of speciation in threatened species recovery. Conservation Letters 2:
the marine environment have both differences and 101–108. (6)
similarities to the terrestrial environment. (2) Brinson, M. M. and S. D. Eckles. 2011. U. S. Department
Bowler, D. E., L. M. Buyung-Ali, J. R. Healey, J. P. G. of Agriculture conservation program and practice
Jones, et al. 2012. Does community forest manage- effects on wetland ecosystem services: A synthesis.
ment provide global environmental benefits and Ecological Applications 21: S116–S127. (10)
improve local welfare? Frontiers in Ecology and the Brodie, J. F., O. E. Helmy, W. Y. Brockelman, and J. L.
Environment 10: 29–36. (11) Maron. 2009. Bushmeat poaching reduces the seed
Boydell, S. and H. Holzknecht. 2003. Land—caught in the dispersal and population growth rate of a mammal-
conflict between custom and commercialism. Land dispersed tree. Ecological Applications 19: 854–863. (8)
Use Policy 20: 203–207. (9) Brodin, T. J. Fick, M. Jonsson, and J. Klaminder. 2013.
Boyles, J. G., P. M. Cryan, G. F. McKracken, and T. H. Dilute concentrations of a psychiatric drug alter
Kunz. 2011. Economic importance of bats in agricul- behavior of fish from natural populations. Science
ture. Science 332: 41–42. (3) 339 (6121): 814–815. (4)
Bradley, B. A., D. M. Blumenthal, R. Early, E. D. Grosholz, Brook, A., M. Zint, and R. DeYoung. 2003. Landowner’s
et al. 2012. Global change, global trade, and the next response to an Endangered Species Act listing and
wave of plant invasions. Frontiers in Ecology and the implications for encouraging conservation. Conser-
Environment 10: 20–28. (4) vation Biology 17: 1638–1649. (5, 6)
Bradshaw, A. D. 1990. The reclamation of derelict land Brooks, J. S. 2016. Design features and project age
and the ecology of ecosystems. In W. R. Jordan III, contribute to joint success in social, ecological, and
M. E. Gilpin, and J. D. Aber (eds.), Restoration Ecol- economic outcomes of community-based conser-
Bibliography 429

vation projects. Conservation Letters. doi: 10.1111/ Burke, L., K. Reytar, M. Spalding, and A. Perry. 2011.
conl.1223. (9) Reefs at Risk Revisited. World Resources Institute,
Brooks, T. M., S. L. Pimm, V. Kapos, and C. Ravilious. Washington DC, USA. (4)
1999. Threat from deforestation to montane and Burkhead, N. M. 2012. Extinction rates in North Ameri-
lowland birds and mammals in insular Southeast can freshwater fishes, 1900–2010. BioScience 62(9):
Asia. Journal of Animal Ecology 68(6): 1061–1078. (5) 798–808. (5)
Brown, M. L., T. M. Donovan, W. Scott-Schwenk, and Burrell, J. 2013. Path of the Pronghorn—Leading to New
D. M. Theobald. 2014. Predicting impacts of future Passages: Part 3. Newswatch, National Geographic.
population growth and development on occupancy newswatch.nationalgeographic.com/2013/12/06/path-of-
rates of forest dependent birds. Biological Conserva- the-pronghorn-leading-to-new-passages-part-3. (8)
tion 170: 311–320. (1) Bussière, E., L. G. Underhill, and R. Altwegg. 2015. Pat-
Brühl, C. A., T. Schmidt, S. Pieper, and A. Alscher. 2013. terns of bird migration phenology in South Africa
Terrestrial pesticide exposure of amphibians: An suggest northern hemisphere climate as the most
underestimated cause of global decline? Scientific consistent driver of change. Global Change Biology
Reports 3, Article 1135. doi: 10.1038/srep01135. (4) 21(6): 2179–2190. (4)
Brundtland, G., M. Khalid, S. Agnelli, S. Al-Athel, et al. BusinessWire. 2015. American Burying Beetle–Mitigation
1987. Our Common Future (“Brundtland report”). Solutions USA Muddy Boggy Endangered Species
United Nations Report of the World Commission on Bank Expansion in Oklahoma. July 30, 2015. www.
Environment and Development. (11) businesswire.com/news/home/20150730006047/en/
Brush, S. B. 2004. Growing biodiversity. Nature 430: American-Burying-Beetle-%E2%80%93-Mitigation-
967–968. (9) Solutions-USA. (11)
Brush, S. B. 2007. Farmers’ rights and protection of tra- Butchart, S. H. M., M. Clarke, R. J. Smith, R. E. Sykes, et
ditional agricultural knowledge. World Development al. 2015. Shortfalls and solutions for meeting na-
35: 1499–1514. (7) tional and global conservation targets. Conservation
Letters 8(5): 329–337. (8, 11)
Bryant, D., L. Burke, J. McManus, and M. Spalding.
1998. Reefs at Risk: A Map-Based Indicator of Threats Butchart, S. H. M., A. J. Stattersfield, and T. M.
to the World’s Coral Reefs. World Resources Institute, Brooks. 2006. Going or gone: definingPossibly
Washington, DC. (4) Extinct’species to give a truer picture of recent
extinctions. Bulletin-British Ornithologists Club 126:
Buchalski, M. R., A. Y. Navarro, W. M. Boyce, T. W.
7. (5)
Vickers, et al. 2015. Genetic population structure of
Peninsular bighorn sheep (Ovis canadensis nelsoni) Butchart, S. H. M., M. Walpole, B. Collen, A. Van Strien,
indicates substantial gene flow across US–Mexico et al. 2010. Global biodiversity: indicators of recent
border. Biological Conservation 184: 218–228. (6) declines. Science 328(5982): 1164–1168. (4)
Buchholz, R. 2007. Behavioral biology: An effective and Butler, R. A., L. P. Koh, and J. Ghazoul. 2009. REDD in
relevant conservation tool. Trends in Ecology and the red: palm oil could undermine carbon payment
Evolution 22: 401–407. (7) schemes. Conservation Letters 2: 67–73. (3)
Buckley, R. 2009. Parks and tourism. PLoS Biology 7:
e1000143. (3) Cafaro, P. 2015. Three ways to think about the sixth mass
Bull, A. T. 2004. Microbial Diversity and Bioprospecting. extinction. Biological Conservation 192: 387–393. (12)
ASM Press, Washington, DC. (3) Caillouet, C. W., B. J. Gallaway, and A. M. Landry. 2015.
Bulman, C. R., R. J. Wilson, A. R. Holt, A. L. Galvez-Bra- Cause and call for modification of the bi-national
vo, et al. 2007. Minimum viable metapopulation size, recovery plan for the Kemp’s ridley sea turtle (Lepi-
extinction debt, and the conservation of declining dochelys kempii)—Second Revision. Marine Turtle
species. Ecological Applications 17: 1460–1473. (5, 6) Newsletter 145: 1–4. (1)
Burgess, M. G., S. Polasky, and D. Tilman. 2013. Predict- Caillouet, C. W., Jr., B. A. Robertson, C. T. Fontaine, T. D.
ing overfishing and extinction threats in multispe- Williams, et al. 1997. Distinguishing captive-reared
cies fisheries. Proceedings of the National Academy of from wild Kemp’s ridleys. Marine Turtle Newsletter
Sciences USA 110: 15943–15948. (4) 77: 1–6. (1)
Burgess, S. C., K. J. Nickols, C. D. Griesmer, L. A. K. Bar- Cain, M. L., W. D. Bowman, and S. D. Hacker. 2014. Ecol-
nett, et al. 2014. Beyond connectivity: How empirical ogy 3rd Ed. Sinauer Associates, Sunderland, MA. (2)
methods can quantify population persistence to Callaway, R. M., G. C. Thelen, A. Rodriguez, and W. E.
improve marine protected-area design. Ecological Holben. 2004. Soil biota and exotic plant invasion.
Applications 24: 257–270. (8) Nature 427(6976): 731–733. (4)
Burghardt, K. T., D. W. Tallamy, and W. G. Shriver. 2009. Calmy, A., E. Goemaere, and G. Van Cutsem. 2015. HIV
Impact of native plants on bird and butterfly biodi- and Ebola virus: two jumped species but not two of
versity in suburban landscapes. Conservation Biology a kind. AIDS 29(13): 1593–1596. (4)
23: 219–224. (10) Camacho, A. E. 2007. Can regulation evolve? Lessons
from a study in maladaptive management. UCLA
Law Review 55: 293–358. (6)
430 Bibliography

Canessa, S., S. J. Converse, M. West, N. Clemann, et al. Ploeotia costata (Euglenozoa) isolated from the
2015. Planning for ex situ conservation in the face Western North Pacific (Taiwan) reveal substantial
of uncertainty. Conservation Biology. doi: 10.1111/ genetic differences. Journal of Eukaryotic Microbiology
cobi.12613 (7) 62: 318–326. (2)
Cannon, J. R. 1996. Whooping crane recovery: a case Charky, N. 2014. Congress attempts to change captiv-
study in public and private cooperation in the ity rules for orcas, marine life. LA Times, Monday,
conservation of endangered species. Conservation March 14, 2014. (12)
Biology10: 813–821. (3) Chen, X. Y. and F. He. 2009. Speciation and endemism
Carey, M. P., B. L. Sanderson, K. A. Barnas, and J. D. under the model of island biogeography. Ecology
Olden. 2012. Native invaders–challenges for science, 90(1): 39–45. (5)
management, policy, and society. Frontiers in Ecology Chen, I., J. K. Hill, R. Ohlemüller, D. B. Roy, and C. D.
and the Environment 10(7): 373–381. (4) Thomas. 2011. Rapid range shifts of species associ-
Carlson, T. 2013. The Politics of a Tree: How a species ated with high levels of climate warming. Science
became national policy. In A. Sher and M. F. Quigley 333: 1024–1026. (4)
(eds.) Tamarix: A Case Study of Ecological Change in Chicago Wilderness. www.chicagowilderness.org (8)
the American West. Oxford University Press. (12) Chittaro, P. M., I. C. Kaplan, A. Keller, and P. S. Levin.
Carnicer, J., M. Coll, M. Ninyerola, X. Pons, et al. 2011. 2010. Trade-offs between species conservation and
Widespread crown condition decline, food web dis- the size of marine protected areas. Conservation Biol-
ruption, and amplified tree mortality with increased ogy 24: 197–206. (5)
climate change-type drought. Proceedings of the Chivian, E. and A. Bernstein (eds.). 2008. Sustaining Life:
National Academy of Sciences USA 108: 1474–1478. (4) How Human Health Depends on Biodiversity. Oxford
Carraro, C., J. Eyckmans, and M. Finus. 2006. Optimal University Press, New York. (3)
transfers and participation decisions in international Christian, C., D. Ainley, M. Bailey, P. Dayton, et al. 2013.
environmental agreements. The Review of Interna- A review of formal objections to Marine Steward-
tional Organizations 1: 379–396. (6) ship Council fisheries certifications. Biological
Carrière, S. M., E. Rodary, P. Méral, G. Serpantié, et al. Conservation 161: 10–17. (4, 12)
2013. Rio +20, biodiversity marginalized. Conserva- Cinner, J. E. and S. Aswani. 2007. Integrating customary
tion 6: 6–11. (9, 11) management into marine conservation. Biological
Carroll, C., R. J. Frederickson, and R. C. Lacy. 2014. De- Conservation 140: 201–216. (4)
veloping metapopulation connectivity criteria from Clare, S. and I. F. Creed. 2014. Tracking wetland loss to
genetic and habitat data to recover the endangered improve evidence-based wetland policy learning
Mexican wolf. Conservation Biology 28: 76–86. (6) and decision making. Wetlands Ecology and Manage-
Carson, R. 1965. The Sense of Wonder. Harper & Row, New ment 22: 235–245. (10)
York. (3) Clark, P. and D. E. Johnson. 2009. Wolf-cattle interactions
Carson, R. 1962. Silent Spring. Houghton Mifflin Com- in the Northern Rocky Mountains. In Range Field
pany. (1) Data 2009 Progress Report. Special Report 1092, pp.
Carter, I., J. Foster, and L. Lock. 2016. The role of animal 1–7. Corvallis, OR: Oregon State University, Agri-
translocations in conserving british wildlife: an cultural Experiment Station. (7)
overview of recent work and prospects for the fu- Clavero, M. and E. García-Berthou. 2005. Invasive species
ture. EcoHealth. doi: 10.1007/s10393-015-1097-1. (7) are a leading cause of animal extinctions. Trends in
Cassimon, D., D. Essers, and A. Fauzi. 2014. Indonesia’s Ecology and Evolution 20: 110–110. (4)
debt-for-development swaps: past, present, and Clayton, S. and G. Myers. 2015. Conservation Psychology:
future. Bulletin of Indonesian Economic Studies 50(1): Understanding and Promoting Human Care for Nature.
75–100. (11) John Wiley & Sons. (11)
Castelletta, M., N. S. Sodhi, and R. Subaraj. 2000. Heavy Clements, T., H. J. Rainey, D. An, V. Rours, et al. 2013. An
extinctions of forest avifauna in Singapore: Lessons evaluation of the effectiveness of a direct payment
for biodiversity conservation in Southeast Asia. for biodiversity conservation: the Bird Nest Protec-
Conservation Biology 14: 1870–1880. (5) tion Program in the northern plains of Cambodia.
Catlin J., M. Hughes, T. Jones, R. Jones, and R. Campbell. Biological Conservation 157: 50–59. (11)
2013. Valuing individual animals through tour- Clewell, A. F. and J. Aronson. 2006. Motivations for the
ism: Science or speculation? Biological Conservation restoration of ecosystems. Conservation Biology 20:
157:93–98. (3) 420–428. (10)
Ceballos, G., P. R. Ehrlich, A. D. Barnosky, A. García, R. Climate Central. 2012. Global Weirdness: Severe Storms,
M. Pringle, and T. M. Palmer. 2015. Accelerated Deadly Heat Waves, Relentless Drought, Rising Seas,
modern human–induced species losses: Entering and the Weather of the Future. Pantheon, New York. (4)
the sixth mass extinction. Science Advances 1(5): CNN. 2016. SeaWorld’s orcas will be last generation at
e1400253. (5) parks. March 16, 2016. (12)
Chan, Y. F., K.-P. Chiang, J. Chang, Ø. Moestrup, and
C.-C. Chung. 2015. Strains of the morphospecies
Bibliography 431

CNN. 2015. Zimbabwean officials: American man versity hotspot impacts mixed-species bird flocks.
wanted in killing of Cecil the lion. July 28, 2015. (3) Biological Conservation 188: 61–71. (4)
Cockle, K. L. and K. Martin. 2015. Temporal dynamics of Cork, S. J., T. W. Clark, and N. Mazur. 2000. Introduction:
a commensal network of cavity-nesting vertebrates: An interdisciplinary effort for koala conservation.
increased diversity during an insect outbreak. Ecol- Conservation Biology 14: 606–609. (8)
ogy 96(4): 1093–1104. (2) Corlett, R. T. 2011. Impacts of warming on tropical low-
Cohen-Shacham, E., T. Dayan, R. de Groot, C., Beltrame, land rainforests. Trends in Ecology and Evolution 26:
et al. 2015. Using the ecosystem services concept to 606–613. (4)
analyse stakeholder involvement in wetland man- Corlett, R. T. 2013. Singapore: half full or half empty? In
agement. Wetlands Ecology and Management 23(2): N. S. Sodhi, L. Gibson, and P. H. Raven (eds.) Con-
241–256. (11) servation Biology: Voices from the Tropics, pp. 142–147.
Colding, J., J. Lundberg, S. Lundberg, and E. Andersson. Wiley-Blackwell, Oxford, UK. (8)
2009. Golf courses and wetland fauna. Ecological Corlett, R. T. 2016. Restoration, reintroduction, and re-
Applications 19: 1481–1491. (9) wilding in a changing world. Trends in Ecology and
Cole, T. 1965. Essay on American Scenery. In J. W. Mc- Evolution. doi: 10.1016/j.tree.2016.017. (10)
Coubrey (ed.), American Art, 1700–1960, pp. 98–109. Corlett, R. T. and R. B. Primack. 2010. Tropical Rain
Prentice-Hall, Englewood Cliffs, NJ. (1) Forests: An Ecological and Biogeographical Comparison,
Cole, I. A., S. M. Prober, I. D. Lunt, and T. B. Koen. 2016. 2nd Ed. Blackwell Publishing, Malden, MA. (2, 3, 4)
A plant traits approach to managing legacy species Corral-Verdugo, V., M. Bonnes, C. Tapia-Fonllem, B. Frai-
during restoration transitions in temperate euca- jo-Sing, et al. 2009. Correlates of pro-sustainability
lypt woodlands. Restoration Ecology. doi: 10.1111/ orientation: The affinity towards diversity. Journal of
rec.12334. (10) Environmental Psychology 29: 34–43. (1)
Colglazier, W. 2015. Sustainable development agenda Costa Rica National Parks. www.costarica-nationalparks.
2030. Science 349:1048–1050. (11) com (11)
Colón-Rivera, R. J., K. Marshall, F. J. Soto-Santiago, Costanza, R., R. de Groot, P. Sutton, S. van der Ploeg, et
D. Ortiz-Torres, and C. E. Flower. 2013. Moving al. 2014. Changes in the global value of ecosystem
forward: Fostering the next generation of Earth services. Global Environmental Change 26: 152–158. (3)
stewards in the STEM disciplines. Frontiers in Ecol- Costello, C., J. M. Drake, and D. M. Lodge. 2007.
ogy and the Environment 11: 383–391. (12) Evaluating an invasive species policy: Ballast water
Colwell, R., S. Avery, J. Berger, G. E. Davis, et al. 2012. exchange in the Great Lakes. Ecological Applications
Revisiting Leopold: Resource Stewardship in the Na- 17: 655–662. (4)
tional Parks. National Park System Advisory Board, Costello, M. J. 2015. Biodiversity: The known, the
Washington, DC. (8) unknown and rates of extinction. Current Biology
Comeau, S., R. C. Carpenter, C. A. Lantz, and P. J. 25(6): R368–371. (2)
Edmunds. 2015. Ocean acidification accelerates Costello, M. J., R. M. May, and N. E. Stork. 2013. Can we
dissolution of experimental coral reef communities. name the Earth’s species before they go extinct? Sci-
Biogeosciences 12(2): 365–372. (4) ence 339: 413–416. (12)
Commission for the Conservation of Antarctic Marine Courchamp, F., J. A. Dunne, Y. Le Maho, R. M. May, et al.
Living Resources (CCAMLR). www.ccamlr.org (6) 2015. Fundamental ecology is fundamental. Trends
Common, M. and S. Stagl. 2005. Ecological Economics: in Ecology and Evolution 30(1): 9–16. (12)
An Introduction. Cambridge University Press, New Cox, P. A. 2001. Pharmacology, biodiversity and. In S. A.
York. (3) Levin (ed.), Encyclopedia of Biodiversity, Vol. 4, pp.
Connor, E. F. and E. D. McCoy. 2001. Species-area relation- 523–536. Academic Press, San Diego, CA. (3)
ships. In S. A. Levin (ed.), Encyclopedia of Biodiversity Cox, P. A. and T. Elmqvist. 1997. Ecocolonialism and
5: 397–412. Academic Press, San Diego, CA. (5) indigenous-controlled rainforest preserves in Sa-
Conservation International and K. J. Caley. 2008. Biologi- moa. Ambio 26: 84–89. (9)
cal diversity in the Mediterranean Basin. In C. J. Cox, R. L. and E. C. Underwood. 2011. The Importance
Cleveland (ed.), Encyclopedia of Earth. Environmental of Conserving Biodiversity Outside of Protected
Information Coalition, National Council for Science Areas in Mediterranean Ecosystems. PLoS ONE 6(1):
and the Environment, Washington, DC. www.eoearth. e14508. (9)
org/article/Biological_diversity_in_the_Mediterranean_
Basin (2) Cox, W. A., F. R. Thompson III, and J. Faaborg. 2012.
Landscape forest cover and edge effects on songbird
Convention on Biological Diversity. www.cbd.int. (11) nest predation vary by nest predator. Landscape Ecol-
Convention on International Trade in Endangered Species ogy 27(5): 659–669. (4)
of Wild Flora and Fauna (CITES). www.cites.org (6) Creech, T. G., C. W. Epps, R. H. Monello, and J. D.
Cordeiro, N. J., L. Borghesio, M. P. Joho, T. J. Monoski, et Wehausen. 2014. Using network theory to prioritize
al. 2015. Forest fragmentation in an African biodi- management in a desert bighorn sheep metapopula-
tion. Landscape Ecology 29(4): 605–619. (6)
432 Bibliography

Crees, J. J., A. C. Collins, P. J. Stephenson, H. M. Mer- arily unique species in mangrove ecosystems. PLoS
edith, et al. 2016. A comparative approach to assess ONE 8(6): e66686. (4)
drivers of success in mammalian conservation Daszak, P., A. A. Cunningham, and A. D. Hyatt. 2000.
recovery programs. Conservation Biology. In press. Emerging infectious diseases of wildlife—threats
doi: 10.1111/cobi.12652. (11) to biodiversity and human health. Science 287:
Christian, C., D. Ainley, M. Bailey, P. Dayton, et al. 2013. 443–449. (4)
A review of formal objections to Marine Steward- Davenport, C. 2015. National approve landmark climate
ship Council fisheries certifications. Biological accord in Paris. New York Times. www.nytimes.
Conservation 161: 10–17. (12) com/2015/12/13/world/europe/climate-change-accord-
Costello, M. J., B. Vanhoorne, and W. Appeltans. 2015. paris.html?_r=0. (11)
Conservation of biodiversity through taxonomy, Davidson, D. J. and J. Andrews. 2013. Not all about con-
data publication, and collaborative infrastructures. sumption. Science 339: 1286–1287. (4)
Conservation Biology 29(4): 1094–1099. (12) Davis, A. M. and A. R. Moore. 2015. Conservation
Crnokrak, P. and D. A. Roff. 1999. Inbreeding depression potential of artificial water bodies for fish commu-
in the wild. Heredity 83 (1999) 260–270. (5) nities on a heavily modified agricultural flood-
Crone, E. E., M. M. Ellis, W. F. Morris, A. Stanley, et al. plain. Aquatic Conservation: Marine and Freshwater
2013. Ability of matrix models to explain the past Ecosystems. (9)
and predict the future of plant populations. Conser- Davis, M. A. 2009. Invasion Biology. Oxford University
vation Biology 27: 968–978. (6) Press, Oxford, UK. (4)
Crosmary, W. G., S. D. Côté, and H. Fritz. 2015. The as- Dawson, J., F. Patel, R. A. Griffiths, and R. P Young. 2016.
sessment of the role of trophy hunting in wildlife Assessing the global zoo response to the amphibian
conservation. Animal Conservation 18(2): 136–137. (3) crisis through 20-year trends in captive collections.
Crowl, T. A., T. O. Crist, R. R. Parmenter, G. Belovsky, Conservation Biology 30(1): 82–91. (7)
and A. E. Lugo. 2008. The spread of invasive spe- De Grammont, P. C. and A. D. Cuarón. 2006. An evalu-
cies and infectious disease as drivers of ecosystem ation of threatened species categorization systems
change. Frontiers in Ecology and the Environment 6: used on the American continent. Conservation Biol-
238–246. (4) ogy 20: 14–27. (6)
Cruickshank, S. S., A. Ozgul, S. Zumbach, and B. R. Deguise, I. E. and J. T. Kerr. 2006. Protected areas and
Schmidt. 2016. Quantifying population declines based prospects for endangered species conservation in
on presence-only records for Red List assessments. Canada. Conservation Biology 20: 48–55. (9)
Conservation Biology. doi: 10.1111/cobi.12688. (6) De la Riva, I. and S. Reichle. 2014. Diversity and Conser-
Cuperus, R., K. J. Canters, H. A. V. de Hars, and D. S. vation of the Amphibians of Bolivia. Herpetological
Friedman. 1999. Guidelines for ecological compen- Monographs 28(1): 46–65. (4)
sation associated with highways. Biological Conserva- Dearborn, D. C. and S. Kark. 2010. Motivations for con-
tion 90: 41–51. (8) serving urban biodiversity. Conservation Biology 4(2):
Czech, B. 2008. Prospects for reconciling the conflict 432–440. (4)
between economic growth and biodiversity con- Denver Post. 2016. Hickenlooper asks EPA for Superfund
servation with technological progress. Conservation fix at leaking mines. Feb. 29, 2016. (4)
Biology 22: 1389–1398. (1)
Dernbach, J. C. and F. Cheever. 2015. Sustainable devel-
opment and its discontents. Transnational Environ-
Dahles, H. 2005. A trip too far: Ecotourism, politics, and mental Law 4(02): 247–287. (11)
exploitation. Development Change 36: 969–971. (3) Devall, B. and G. Sessions. 1985. Deep Ecology. Perigrine
Danielson, F., P. M. Jensen, N. D. Burgess, R. Altamirano, Smith, Salt Lake, UT. (3)
et al. 2014. A multicountry assessment of tropical Diamond, J. 2005. Collapse: How Societies Choose to Fail or
resource monitoring by local communities. BioSci- Succeed. Penguin Books, New York. (3)
ence 64: 236–251. (8)
Dickman, A. J., E. A. Macdonald, and D. W. Macdonald.
Danielson, F., K. Pirhofer-Walzi, T. P. Adrian, D. R. Kapi- 2011. A review of financial instruments to pay for
jimpanga, et al. 2014. Linking public participation predator conservation and encourage human-
in scientific research to the indicators and needs of carnivore coexistance. Proceedings Of The National
international environmental agreements. Conserva- Academy Of Sciences USA 108: 13937–13944. (9)
tion Letters 7: 12–24. (12)
Dieter, C. D. and D. J. Schaible. 2014. Distribution and
Daru B. H. and P. C. le Roux 2016. Marine protected areas Population Density of Jackrabbits in South Dakota.
are insufficient to conserve global marine plant Great Plains Research 24(2): 127–134. (6)
diversity. Global Ecology and Biogeography 25(3):
324–334. (9) Di Marco, M., B. Collen, C. Rondinini, and G. M. Mace.
2015. Historical drivers of extinction risk: using past
Daru, B. H., K. Yessoufou, L. T. Mankga, and T. J. Davies. evidence to direct future monitoring. Proceedings of
2013. A global trend towards the loss of evolution- the Royal Society B 282(1813): 20150928. (5)
Bibliography 433

Di Minin, E., D. C. Macmillan, P. S. Goodman, B. Escott, Drus, G. M., T. L. Dudley, M. L. Brooks, and J. R. Match-
et al. 2013. Conservation business and conservation ett. 2013. The effect of leaf beetle herbivory on the
planning in a biological diversity hotspot. Conserva- fire behaviour of tamarisk (Tamarix ramosissima
tion Biology 27: 808–820. (11) Lebed.). International Journal of Wildland Fire 22(4):
Di Minin E., N. Leader-Williams, and C. J. A. Bradshaw. 446–458. (4)
2016. Banning trophy hunting will exacerbate bio- Duarte, A., J. S. Hatfield, T. M. Swannack, M. R. Forstner,
diversity loss. Trends in Ecology and Evolution 31(2): et al. 2016. Simulating range-wide population and
99–102. (9) breeding habitat dynamics for an endangered wood-
Dinerstein, E., K. Varma, E. Wikramanayake, G. Powell, land warbler in the face of uncertainty. Ecological
et al. (2013). Enhancing conservation, ecosystem Modelling 320: 52–61. (6)
services, and local livelihoods through a wildlife Duarte, C. M., K. A. Pitt, C. H. Lucas, J. E. Purcell, et.
premium mechanism. Conservation Biology 27(1): al. 2013. Is global ocean sprawl a cause of jellyfish
14–23. (9) blooms? Frontiers in Ecology and the Environment. 11:
Dinerstein, E., K. Varma, E. Wikramanayake, G. Powell, et 91–97. (4)
al. 2013. Enhancing conservation, ecosystem servic- Duchelle, A. E., M. R. Guariguata, G. Less, M. A. Albor-
es, and local livelihoods through a wildlife premium noz, et al. 2012. Evaluating the opportunities and
mechanism. Conservation Biology 27: 14–23. (9) limitations to multiple use of Brazil nuts and timber
Dobson, A. 2005. Monitoring global rates of biodiver- in Western Amazonia. Forest Ecology and Manage-
sity change: Challenges that arise in meeting the ment 268: 39–48. (9)
Convention on Biological Diversity (CBD) 2010 Dudley, N. (ed.) 2008. Guidelines for Applying Protected
goals. Philosophical Transactions of the Royal Society Area Management Categories. IUCN, Gland, Switzer-
B-Biological Sciences 360: 229–241. (3) land. (8)
Dodds, W. K., K. C. Wilson, R. L. Rehmeier, G. L. Knight, Dudley, N., L. Higgins-Zogib, and S. Mansourian. 2009.
et al. 2008. Comparing ecosystem goods and servic- The links between protected areas, faiths, and sa-
es provided by restored and native lands. BioScience cred natural sites. Conservation Biology 23: 568–577.
58: 837–845. (10) (1)
Dolman, P. M., N. J. Collar, K. M. Scotland, and R. J. Dukes, J. S., N. R. Chiariello, S. R. Loarie, and C. B. Field.
Burnside. 2015. Ark or park: stochastic popula- 2011. Strong response of an invasive plant species
tion modeling to evaluate potential effectiveness (Centaurea solstitialis L.) to global environmental
of in situ and ex situ conservation for a critically changes. Ecological Applications 21: 1887–1894. (4)
endangered bustard. Journal of Applied Ecology 52: Dullinger, S., F. Essl, W. Rabitsch, K. Erb, et al. 2013.
841–850. (7) Europe’s other debt crisis caused by the long legacy
Domisch, S., S. C. Jahnig, J. P. Simaika, M. Kuemmerlen, of future extinctions. Proceedings of the National
and S. Stoll. 2015. Application of species distribu- Academy of Sciences USA 110: 7342–7347. (5)
tion models in stream ecosystems: the challenges of Dunn, R. R. 2005. Modern insect extinctions, the neglect-
spatial and temporal scale, environmental predic- ed majority. Conservation Biology 19: 1030–1036. (6)
tors and species occurrence data. Fundamental Ap- Dvořák, P., A. Poulíčková, P. Hašler, M. Belli M., et al.
plied Limnology 186(1–2): 45–61. (2) 2015. Species concepts and speciation factors in
Donovan, G. H., D. T. Butry, Y. L. Michael, J. P. Preste- cyanobacteria, with connection to the problems of
mon, et al. 2013. The relationship between trees diversity and classification. Biodiversity and Conser-
and human health: evidence from the spread of the vation. doi: 10.1007/s10531-015-0888-6. (2)
emerald ash borer. American Journal of Preventive
Medicine 44: 139–145. (3)
Earnst, S. L., D. S. Dobkin, and J. A. Ballard. 2012.
Donlan, J., H. W. Greene, J. Berger, C. E. Bock, et al. 2006. Changes in avian and plant communities of aspen
Re-wilding North America. Nature 436: 913–914. (7) woodlands over 12 years after livestock removal in
Doney, S. C., V. J. Fabry, R. A. Feely, and J. A. Kleypas. the northwestern great basin. Conservation Biology
2009. Ocean acidification: the other CO2 problem. 26: 862–871. (9)
Annual Review of Marine Science 1: 169–92. (4) Ebbin, S. 2009. Institutional and ethical dimensions of
Doughty, C. E. 2013. Preindustrial human impacts on resilience in fishing systems: Perspectives from co-
global and regional environment. Annual Review of managed fisheries in the Pacific Northwest. Marine
Environment and Resources 38: 503–527. (4) Policy 33: 264–270. (1)
Douglas, L. R. and K. Alie. 2014. High-value natu- Economist. 2013. Special Report: Biodiversity. The Econo-
ral resources: Linking wildlife conservation to mist September 14, 2013. (10)
international conflict, insecurity, and development Ecosystem Marketplace. 2010. The Katoomba Group.
concerns. Biological Conservation 171: 270–277. (6) www.ecosystemmarketplace.com (9)
Drayton, B. and R. B. Primack. 1999. Experimental extinc- Edwards, D. P. and S. G. Laurance. 2012. Green label-
tion of garlic mustard (Alliaria petiolata) populations: ing, sustainability and the expansion of tropical
Implications for weed science and conservation agriculture: critical issues for certification schemes.
biology. Biological Invasions 1: 159–167. (5) Biological Conservation 151: 60–64. (12)
434 Bibliography

Ehrenfeld, D. W. 1970. Biological Conservation. Holt, Rine- Fairtrade International. www.fairtrade.net (12)
hart and Winston, New York. (5) Faith, D. P. 2008. Threatened species and the potential
Ehrlich, P. R. and A. H. Ehrlich. 1982. Extinction: The loss of phylogenetic diversity: conservation sce-
Causes and Consequences of the Disappearance of Spe- narios based on estimated extinction probabilities
cies. Gollancz, London. (3) and phylogenetic risk analysis. Conservation Biology
Ehrlich, P. R., P. M. Kareiva, and G. C. Daily. 2012. Secur- 22(6): 1461–1470. (6)
ing natural capital and expanding equity to rescale Falk, D. A., M. A. Palmer, and J. B. Zedler (eds.). 2006.
civilization. Nature 486(7401): 68–73. (4) Foundations of Restoration Ecology: The Science and
Ehrlich, P. R. and R. M. Pringle. 2008. Where does bio- Practice of Ecological Restoration. Island Press, Wash-
diversity go from here? A grim business-as-usual ington, DC. (10)
forecast and a hopeful portfolio of partial solutions. Fanshawe, S., G. R. VanBlaricom, and A. A. Shelly. 2003.
Proceedings of the National Academy of Sciences USA Restored top carnivores as detriments to the perfor-
105: 11579–11586. (10) mance of marine protected areas intended for fish-
Elkinton, J. S., D. Parry, and G. H. Boettner. 2006. Impli- ery sustainability: a case study with red abalones
cating an introduced generalist parasitoid in the and sea otters. Conservation Biology 17: 273–283. (7)
invasive browntail moth’s enigmatic demise. Ecol- Farmer, J. R., D. Knapp, V. J. Meretsky, C. Chancellor,
ogy 87: 2664–2672. (4) and B. C. Fischer. 2011. Motivations influencing the
Elliott, J. E. and K. H. Elliott. 2013. Tracking marine pol- adoption of conservation easements. Conservation
lution. Science 340: 556–558. (4) Biology 25: 827–834. (11)
Elrich, P. R. and J. P. Holdren. 1971. Impact of Population Farmer, J. R., Z. Ma, M. Drescher, E. G. Knackmuhs, and
Growth. Science 171:1212–1217. (4) S. L. Dickinson. 2016. Private landowners, volun-
Emerson, R. W. 1836. Nature. James Monroe and Co., tary conservation programs, and implementation of
Boston. (1) conservation friendly land management practices.
Conservation Letters. (9)
Emerton, L. 1999. Balancing the opportunity costs of
wildlife conservation for communities around Lake Fehr-Duda, H. and E. Fehr. 2016. Sustainability: Game
Mburo National Park, Uganda. Evaluating Eden Se- human nature. Nature 530. (11)
ries Discussion Paper No. 5, International Institute Feist, B. E., E. R. Buhle, P. Arnold, J. W. Davis, and N.
for Environment and Development, London. (3) L. Scholz. 2011. Landscape Ecotoxicology of Coho
Encyclopedia of Life. eol.org (2, 12) Salmon Spawner Mortality in Urban Streams. PLoS
ONE 6(8): e23424. (4)
Encyclopedia of the Nations. 2009. India. www.nationsen-
cyclopedia.com/economies/Asia-and-the-Pacific/India. Felson, A. J., M. A. Bradford, and T. M. Terway. 2013.
html. (4) Promoting Earth Stewardship through urban design
experiments. Frontiers in Ecology and the Environment
Engler, M. 2008. Value of the international trade in wild- 11: 362–367. (10)
life. TRAFFIC Bulletin 22(1): 4–5. (3)
Feng, K., K. Hubacek, and D. Guan. 2009. Lifestyles,
Equator Initiative. www.equatorinitiative.org (9) technology and CO2 emissions in China: A regional
Estes, J. A., J. Terbough, J. S. Brashares, M. E. Power, et al. comparative analysis. Ecological Economics 69(1):
2011. Trophic downgrading of planet earth. Science 145–154. (4)
333: 301–306. (2) Ferrer. M., I. Newton, and M. Pandolfi. 2009. Small popu-
European Commission. 2008. The EU Biodiversity lations and offspring sex-ratio deviations in eagles.
Strategy to 2020. ec.europa.eu/environment/nature/info/ Conservation Biology 23: 1017–1025. (5)
pubs/docs/brochures/2020%20Biod%20brochure%20 Ferro, P. J., M. M. Hanauer, and K. R. E. Sims. 2011.
final%20lowres.pdf (10) Conditions associated with protected area success
Evans, S. R. and B. C. Sheldon. 2008. Interspecific pat- in conservation and poverty reduction. Proceed-
terns of genetic diversity in birds: Correlations with ings of the National Academy of Sciences USA 108:
extinction risk. Conservation Biology 22: 1016–1025. 13913–13918. (8)
(5) Fischer, J. and D. B. Lindenmayer. 2000. An assessment
Ewing, S. R., R. G. Nager, M. A. C. Nicoll, A. Aumjaud, of published results of animal relocations. Biological
et al. 2008. Inbreeding and loss of genetic variation Conservation 96: 1–11. (7)
in a reintroduced population of Mauritius kestrel. Fischer, J., R. Dyball, I. Fazey, C. Gross, et al. 2012. Hu-
Conservation Biology 22: 395–404. (5) man behavior and sustainability. Frontiers in Ecology
and the Environment 10: 153–160. (12)
Fadeeva, N., V. Mordukhovich, and J. Zograf. 2015. New Fischer, L. K., V. Rodorff, M. von der Lippe, and I.
deep-sea large free-living nematodes from maroben- Kowarik. 2016. Drivers of biodiversity patterns in
thos in the Kuril-Kamchatka trench (the North-West parks of a growing South American megacity. Urban
Pacific). Deep-Sea Research II 111:95–103. (2) Ecosystems 1–19. (9)
Faeth, S. H., P. S. Warren, E. Shochat, and W. A. Marus- Fisher, D. O. and S. P. Blomberg. 2012. Inferring extinc-
sich. 2005. Trophic dynamics in urban communities. tion of mammals from sighting records, threats, and
BioScience 55: 399–407. (2) biological traits. Conservation Biology 26: 57–67. (5)
Bibliography 435

Fisher, M. 2016. Fall, resurrection and uncertainty: an Linking knowledge to action. Conservation Letters 5:
Arabian tale. Oryx 50(01): 1–2. (7) 1–10. (8)
Fisher, R., R. A. O’Leary, S. Low-Choy, K. Mengersen, K., Fragkias, M., J. Lobo, D. Strumsky, and K. C. Seto. 2013.
et al. 2015. Species Richness on Coral Reefs and the Does size matter? Scaling of CO2 emissions and US
pursuit of convergent global estimates. Current Biol- urban areas. PLoS ONE 8(6): e64727. doi: 10.1371/
ogy 25: 500–505. (2) journal.pone.0064727. (4)
Fisk, M. R., S. J. Giovannoni, and I. H. Thorseth. 1998. Al- Frankham R. 1995. Effective population size adult
teration of oceanic volcanic glass: textural evidence population size ratios in wildlife – a review. Genetics
of microbial activity. Science 281: 978–980. (2) Research 166: 95–107. (5)
Flather, C. H., G. D. Hayward, S. R. Beissinger, and P. Frankham, R. 2005. Resolving the genetic paradox in
A. Stephens. 2011. Minimum viable populations: Is invasive species. Heredity 94(4): 385–385. (5)
there a “magic number” for conservation practitio- Frankham, R., C. J. A. Bradshaw, and B. W. Brook. 2014.
ners? Trends in Ecology and Evolution 26: 307–316. (6) Genetics in conservation management: revised
Flinn, K. M. and P. L. Marks. 2007 Agricultural legacies in recommendations for the 50/500 rules, Red List
forest environments: tree communities, soil proper- criteria and population viability analyses. Biological
ties, and light availability. Ecological Applications 17: Conservation 170: 56–63. (5)
452–463. (10) Frankham, R. 2005. Genetics and extinction. Biological
Flohre, A., C. Fischer, T. Aavik, J. Bengtsson, et al. 2011. Conservation 126: 131–140. (5)
Agricultural intensification and biodiversity Frankham, R., J. D. Ballou, and D. A. Briscoe. 2009. Intro-
partitioning in European landscapes comparing duction to Conservation Genetics, 2nd Ed. Cambridge
plants, carabids, and birds. Ecological Applications 21: University Press, Cambridge, UK. (2, 5)
1772–1780. (2) Frankham, R., J. D. Ballou, M. R. Dudash, M. D. B.
Flory, S. L. and K. Clay. 2009. Effects of roads and forest Eldridge, et al. 2012. Implications of different spe-
successional age on experimental plant invasions. cies concepts for conserving biodiversity. Biological
Biological Conservation. 142: 2531–2537. (4) Conservation 153: 25–31. (2)
Fontaine, B., P. Bouchet, K. Van Achterberg, M. A. Fraser, L. H., W. L. Harrower, H. W. Garris, S. Davidson,
Alonso-Zarazaga, et al. 2007. The European Union’s et al. 2015. A call for applying trophic structure
2010 target: Putting rare species in focus. Biological in ecological restoration. Restoration Ecology 23:
Conservation 139: 167–185. (6) 503–507. (10)
Food and Agriculture Organization of the United Na-
tions (FAO). 2007. FAOSTAT. faostat.fao.org (7) Gagic, V., I. Bartomeus, T. Jonsson, A. Taylor, et al. 2015
Ford, A. T., A. P. Clevenger, and A. Bennett. 2009. Com- Functional identity and diversity of animals predict
parison of methods of monitoring wildlife crossing ecosystem functioning better than species-based indi-
structures on highways. Journal of Wildlife Manage- ces. Proceedings of the Royal Society B 282: 20142620. (2)
ment 73: 1213–1222. (8) Gagné, S. A., F. Eigenbrod, D. G. Bert, G. M. Cunnington,
Forister, M. L., A. C. McCall, N. J. Sanders, J. A. Fordyce, et al. 2015. A simple landscape design framework
et al. 2010. Compounded effects of climate change for biodiversity conservation. Landscape and Urban
and habitat alteration shift patterns of butterfly di- Planning 136: 13–27. (8)
versity. Proceedings of the National Academy of Sciences Gaines, S. D., C. White, M. H. Carr, and S. R. Palumbi.
USA 107: 2088–2092. (4) 2010. Designing marine reserve networks for both
Forsman, A. 2014. Effects of genotypic and phenotypic conservation and fisheries management. Proceed-
variation on establishment are important for conser- ings of the National Academy of Sciences USA 107:
vation, invasion, and infection biology. Proceedings 18286–18293. (8)
of the National Academy of Sciences 111(1): 302–307. (5) Galbraith, C. A., P. V. Grice, G. P. Mudge, S. Parr, and M.
Forzza, R. C., J. F. A. Baumgratz, C. E. M. Bicudo, D. A. W. Pienkowski. 1998. The role of statutory bodies in
L. Canhos, et al. 2012. New Brazilian floristic list ornithological conservation. Ibis 137: S224–S231. (1)
highlights conservation challenges. Bioscience 62: Galatowitsch, S. M. Ecological Restoration. 2012. Sinauer
39–45. (2) Associates, Sunderland, MA. (10)
Foster, B. L., K. Kindscher, G. R. Houseman, and C. A. Galetti, M., and R. Dirzo. 2013. Ecological and evolution-
Murphy. 2009. Effects of hay management and na- ary consequences of living in a defaunated world.
tive species sowing on grassland community struc- Biological Conservation 163: 1–6. (4)
ture, biomass, and restoration. Ecological Applications
19: 1884–1896. (10) Gallardo, B., M. Clavero, M. I. Sánchez, and M. Vilá.
2016. Global ecological impacts of invasive species
Foster, K. R., P. Vecchia, and M. H. Repacholi. 2000. Sci- in aquatic ecosystems. Global Change Biology. doi:
ence and the precautionary principle. Science 288: 10.1111/gcb.13004. (4)
979–981. (3)
Gardiner, S., S. Caney, D. Jamieson, and H. Shue. 2010.
Fox, H. E., M. B. Mascia, X. Basurto, A. Costa, et al. 2012. Climate Ethics: Essential Readings. Oxford University
Reexamining the science of marine protected areas: Press, New York. (3)
436 Bibliography

Garibaldi, L. A., I. Steffan-Dewenter, R. Winfree, M. A. Scarcity and Diversity, pp. 19–34. Sinauer Associates,
Aizen, et al. 2013. Wild pollinators enhance fruit set Sunderland, MA. (5)
of crops regardless of honey bee abundance. Science Gilstad-Hayden, K., L. R. Wallace, A. Carroll-Scott, S.
339(6127): 1608–1611. (3) R. Meyer, et al. 2015. Research note: Greater tree
Garrard, G. E., F. Fidler, B. C. Wintle, Y. E. Chee, and S. A. canopy cover is associated with lower rates of both
Bekessy. 2015. Beyond advocacy: making space for violent and property crime in New Haven, CT.
conservation scientists in public debate. Conserva- Landscape and Urban Planning 143: 248–253. (3)
tion Letters. doi: 10.1111/conl.12193. (12) Gliessman, S. R. 2015. Agroecology: The Ecology of Sustain-
Garzón-Machado, V., J. M. González-Mancebo, A. able Food Systems. CRC Press, Boca Raton, FL. (7)
Palomares-Martínez, A. Acevedo-Rodríguez, et al. Global Environment Facility (GEF): Investing in Our
2010. Strong negative effect of alien herbivores on Planet. www.thegef.org/gef. (11)
endemic legumes of the Canary pine forest. Biologi- Global Footprint Network: Advancing the Science of
cal Conservation 143: 2685–2694. (4) Sustainability. 2006. www.footprint.org (4)
Gaston, K. J. 2010. Urban Ecology. Cambridge University Godefroid, S., C. Piazza, G. Rossi, S. Buord, et al. 2011a.
Press, New York. (4) How successful are plant species reintroductions?
Gaston, K. J. and J. I. Spicer. 2004. Biodiversity: An Introduc- Biological Conservation 144: 672–682. (7)
tion, 2nd Ed. Blackwell Publishing, Oxford, UK. (2) Godefroid, S., S. Rivière, S. Waldren, N. Boretos, R. East-
Gates, C. C., P. Jones, M. Suitor, A. Jakes, et al. 2012. The wood, et al. 2011b. To what extent are threatened
influence of land use and fences on habitat effective- European plant species conserved in seed banks?
ness, movements and distribution of pronghorn in Biological Conservation 144: 1494–1498. (7)
the grasslands of North America. In M. J. Somers Godfray, H. C. J., J. R. Beddington, I. R. Crute, L. Had-
and M. W. Hayward (eds.), Fencing for Conservation, dad, et al. 2010. Food security: The challenge of
pp. 277–294. Springer, New York. (4) feeding 9 billion people. Science 327: 812–818. (4)
Gavin, M. C., J. McCarter, A. Mead, F. Berkes, et al. 2015. Goetz, S. J., M. Sun, S. Zolkos, A. Hansen, and R.
Defining biocultural approaches to conservation. Dubayah. 2014. The relative importance of climate
Trends in Ecology and Evolution 30(3): 140–145. (9) and vegetation properties on patterns of North
Geldmann, J., M. Barnes, L. Coad, I. D. Craigie, et al. American breeding bird species richness. Environ-
2013. Effectiveness of terrestrial protected area mental Research Letters 9(3): 034013. (6)
in reducing habitat loss and population declines. González E., A. A. Sher, E. Tabacchi, A. Masip, and M.
Biological Conservation 161: 230–238. (8) Poulin. 2015. Restoration of riparian vegetation: A
Gerrodette, T. and W. G. Gilmartin. 1990. Demographic global review of implementation and evaluation
consequences of changing pupping and hauling approaches in the international, peer-reviewed
sites of the Hawaiian monk seal. Conservation Biol- literature. Journal of Environmental Management 158:
ogy 4: 423–430. (6) 85–94. (10)
Gessner, M. O., C. M. Swan, C. I. Dang, B. G. McKie, et González-Maya, J. F., L. R. Víquez-R, A. Arias-Alzate, J.
al. 2010. Diversity meets decomposition. Trends in L. Belant, and G. Ceballos. 2016. Spatial patterns of
Ecology and Evolution 25: 372–380. (2) species richness and functional diversity in Costa
Gibbs, K. E. and D. J. Currie, D. J. 2012. Protecting endan- Rican terrestrial mammals: implications for conser-
gered species: do the main legislative tools work? vation. Diversity and Distributions 22: 43–56. (2)
PLoS One 7(5): e35730. (6) Gore, A. 2006. An Inconvenient Truth: The Planetary Emer-
Gibson, L., A. J. Lynam, C. J. A. Bradshaw, F. He, et al. gency of Global Warming and What We Can Do About
2013. Near-complete extinction of native small It. Rodale Books, New York. (4)
mammal fauna 25 years after forest fragmentation. Goss, J. R. and G. S. Cumming. 2013. Networks of
Science 341: 1508–1510. (4) wildlife translocations in developing countries: an
Gibson, L., T. M. Lee, L. P. Koh, B. W. Brook, et al. 2011. emerging conservation issue? Frontiers in Ecology
Primary forests are irreplaceable for sustaining and the Environment 11(5): 243–250. (4)
tropical biodiversity. Nature 478: 378–381. (4) Gotelli, N. J., A. Chao, R. K. Colwell, W. H. Hwang, and
Gilbert-Norton, L., R. Wilson, J. R. Stevens, and K. H. G. R. Graves. 2012. Specimen-based modeling,
Beard. 2010. A meta-analytic review of corridor ef- stopping rules, and the extinction of the ivory-billed
fectiveness. Conservation Biology 24: 660–668. (8) woodpecker. Conservation Biology 26(1): 47–56. (5)
Gilmour, J. P., L. D. Smith, A. J. Heyward, A. H. Baird, Granek, E. F., E. M. P. Madin, M. A. Brown, W. Figueira,
and M. S. Pratchett. 2013. Recovery of an isolated et al. 2008. Engaging recreational fishers in manage-
coral reef system following severe disturbance. Sci- ment and conservation: Global case studies. Conser-
ence 340: 69–71. (4) vation Biology 22: 1125–1134. (12)
Gilpin, M. E. and M. E. Soulé. 1986. Minimum viable Grant, B. R. and P. R. Grant. 2008. Fission and fusion of
populations: Processes of species extinction. In M. Darwin’s finches populations. Philosophical Transac-
E. Soulé (ed.), Conservation Biology: The Science of tions of the Royal Society. Series B: Biological Sciences
363: 2821–2829. (5)
Bibliography 437

Grassle, J. F. 2001. Marine ecosystems. In S. A. Levin re-introduce endangered wild dogs in South Africa.
(ed.), Encyclopedia of Biodiversity, Vol. 4, pp. 13–26. Journal of Applied Ecology 45: 100–108. (7)
Academic Press, San Diego, CA. (2) Gustafsson, L., A. Felton, A. M. Felton, J. Brunet, et al.
Gray, L. K., T. Gylander, M. S. Mbogga, P. Chen, A. Ha- 2015. Natural versus national boundaries: the im-
mann. 2011. Assisted migration to address climate portance of considering biogeographical patterns in
change: Recommendations for aspen reforesta- forest conservation policy. Conservation Letters 8(1):
tion in western Canada. Ecological Applications 21: 50–57. (12)
1591–1603. (7) Gutiérrez, J. L. and C. Bernstein. 2014. Ecosystem impacts
Greene, R. M., J. C. Lehrter, and J. D. Hagy. 2009. Mul- of invasive species. BIOLIEF 2011 2nd World Con-
tiple regression models for hindcasting and fore- ference on Biological Invasion and Ecosystem Func-
casting midsummer hypoxia in the Gulf of Mexico. tioning, Mar del Plata, Argentina, 21–24 November
Ecological Applications 19: 1161–1175. (4) 2011. Acta Oecologica 54: 1–138. (4)
Grenier, M. B., D. B. McDonald, and S. W. Buskirk. 2007.
Rapid population growth of a critically endangered Hall, J. A. and E. Fleishman. 2010. Demonstration as a
carnivore. Science 317: 779. (7) means to translate conservation science into prac-
Griffiths, R. A. and L. Pavajeau. 2008. Captive breeding, tice. Conservation Biology 24: 120–127. (1)
reintroduction, and the conservation of amphibians. Halpern, B. S. 2014. Making marine protected areas
Conservation Biology 22: 852–861. (7) work. Nature 506: 167–168. (1)
Grman, E., T. Bassett, and L. A. Brudvig. 2014. A prairie Halpern, B. S., C. J. Klein, C. J. Brown, M. Beger, et al.
plant community data set for addressing questions 2013. Achieving the triple bottom line in the face of
in community assembly and restoration. Ecology 95: inherent trade-offs among social equity, economic
2363. (10) return, and conservation. Proceedings of the National
Groom, M. J., G. K. Meffe, and C. R. Carroll (eds.). 2006. Academy of Sciences USA 110: 6229–6234. (12)
Principles of Conservation Biology, 3rd Ed. Sinauer Halpern, B. S., K. A. Selkoe, F. Micheli, and C. V. Kappel.
Associates, Sunderland, MA. (2, 3, 4, 5) 2007. Evaluating and ranking the vulnerability of
Groombridge, B. and M. D. Jenkins. 2010. World Atlas of global marine ecosystems to anthropogenic threats.
Biodiversity; Earth’s living resources in the 21st century. Conservation Biology 21: 1301–1315. (10)
University of California Press, Berkeley. (2) Halsey, S. J., T. J. Bell, K. McEachern, and N. B. Pavlovic.
Gross, L. 2008. Can farmed and wild salmon coexist? 2015. Comparison of reintroduction and enhance-
PLoS Biology 6: e46. (3) ment effects on metapopulation viability. Restoration
Groves, C., E. Game, M. Anderson, M. Cross, et al. 2012. Ecology 23(4): 375–384. (7)
Incorporating climate change into systematic con- Hallwass, G., P. F. Lopes, A. A. Juras, and R. A. M. Silva-
servation planning. Biodiversity and Conservation 21: no. 2013. Fishers’ knowledge identifies environmen-
1651–1671. (8) tal changes in fish abundance trends in impounded
Grumbine, R. E. 2007. China’s emergences and the tropical rivers. Ecological Applications 23: 392–407. (8)
prospects for global sustainability. BioScience 57: Hambler, C., P. A. Henderson, and M. R. Speight. 2011.
249–255. (4) Extinction rates, extinction-prone habitats, and indi-
Grumbine, R. E. and J. Xu. 2011. Creating a “Conserva- cator groups in Britain and at larger scales. Biological
tion with Chinese Characteristics.” Biological Conser- Conservation 144: 713–721. (4, 5)
vation 144: 1347–1355. (8) Hammond, P. S., K. Macleod, P. Berggren, D. L. Borchers,
Guerrant, E. O. 1992. Genetic and demographic consid- et al. 2013. Cetacean abundance and distribution in
erations in the sampling and reintroduction of rare European Atlantic shelf waters to inform conserva-
plants. In P. L. Fiedler and S. K. Jain (eds.), Conserva- tion and management. Biological Conservation 164:
tion Biology: The Theory and Practice of Nature Con- 107–122. (6)
servation, Preservation and Management, pp. 321–344. Handel, S. N. 2016. Greens and Greening: Agriculture
Chapman and Hall, New York. (5) and Restoration Ecology in the City. Ecological Resto-
Guerrant, E. O. Jr., K. Havens, and M. Maunder (eds.). ration 34(1): 1–2. (10)
2013. Ex Situ Conservation: Supporting Species Sur- Hanley, N., L. P. Dupuy, and E. McLaughlin. 2015. Genu-
vival in the Wild. Island Press, Washington, DC. (7) ine savings and sustainability. Journal of Economic
Guiry, M. D. 2012 How many species of Algae are there? Surveys 29(4): 779–806. (3)
Journal of Phycology 48: 1057–1063. (2) Hanna, E. and M. Cardillo. 2013. Island mammal extinc-
Gupta, N., R. Raghavan, K. Sivakumar, V. Mathur, and tions are determined by interactive effects of life
A. C. Pinder. 2015. Assessing recreational fisheries history, island biogeography and mesopredator
in an emerging economy: Knowledge, perceptions suppression. Global Ecology and Biogeography 23(4)
and attitudes of catch-and-release anglers in India. 395–404. doi: 10.1111/geb.12103. (5)
Fisheries Research 165: 79–84. (12) Hannah, L., P. R. Roehrdanz, M. Ikegami, A. V. Shepard,
Gusset, M., S. J. Ryan, M. Hofmeyr, G. V. Dyk, et al. 2008. et al. 2013. Climate change, wine, and conservation.
Efforts going to the dogs? Evaluating attempts to Proceedings of the National Academy of Sciences USA
110: 6907–6912. (4)
438 Bibliography

Hansen, A. J., C. R. Davis, N. Piekielek, J. Gross, et planus, and other emerging threats on populations
al. 2011. Delineating the ecosystems containing of the federally threatened Pitcher’s thistle, Cirsium
protected areas for monitoring and management. pitcheri. Biological Conservation 158: 202–211. (4)
BioScience 61: 363–373. (8) Haw, J. 2013. Southern California and endangered aba-
Hansen, M. C., P. V. Potapov, R. Moore, M. Hancher, et al. lone populations. Scientific American, July 3, 2013. (5)
2013. High-resolution global maps of 21st-century Hayes, T. B., P. Falso, S. Gallipeau, and M. Stice. 2010.
forest cover change. Science 342(6160): 850–853. (4) The cause of global amphibian declines: a develop-
Hansen, M. C., S. V. Stehman, P. V. Potapov, B. Arunar- mental endocrinologist’s perspective. The Journal of
wati, et al. 2009. Quantifying changes in the rates of Experimental Biology 213(6): 921–933. (4)
forest clearing in Indonesia from 1990 to 2005 using Hayward, M. W. 2009. Conservation management for the
remotely sensed data sets. Environmental Research past, present, and future. Biodiversity and Conserva-
Letters 4: 034001. (4, 6) tion 18: 765–775. (10)
Hanson, T., T. M. Brooks, G. A. B. da Fonseca, M. Hoff- He, X., M. L. Johansson, and D. D. Heath, D. D. 2016.
mann, et al. 2009. Warfare in biodiversity hotspots. Role of genomics and transcriptomics in selection
Conservation Biology 23: 578–587. (6) of reintroduction source populations. Conservation
Hanski, I., A. Moilanen, and M. Gyllenberg. 1996. Mini- Biology. doi: 10.1111/cobi.12674. (7)
mum viable metapopulation size. American Natural- Heber, S. and J. V. Briskie. 2010. Population bottlenecks
ist 384: 527–541. (5) and increased hatching failure in endangered birds.
Hardin, G. 1968. The tragedy of the commons. Science Conservation Biology 24: 1674–1678. (5)
162: 1243–1248. (3) Hedges, S., A. Johnson, M. Ahlering, M. Tyson, and L. S.
Harding, G., R. A. Griffiths, and L. Pavajeau. 2015. Eggert. 2013. Accuracy, precision, and cost-effective-
Developments in amphibian captive breeding and ness of conventional dung density and fecal DNA
reintroduction programs. Conservation Biology. doi: based survey methods to estimate Asian elephant
10.1111/cobi.12612. (7) (Elephas maximus) population size and structure.
Hardwick, K. A., P. Fiedler, L. C. Lee, B. Pavlik, et al. Biological Conservation 159: 101–108. (6)
2011. The role of botanic gardens in the science and Hedges, S., M. J. Tyson, A. F. Sitompul, M. F. Kinnaird,
practice of ecological restoration. Conservation Biol- et al. 2005. Distribution, status, and conservation
ogy 25: 265–275. (7) needs of Asian elephants (Elephas maximus) in
Hardy, M. J., J. A. Fitzsimons, S. A. Bekessy, and A. Gor- Lampung Province, Sumatra, Indonesia. Biological
don. 2016. Exploring the permanence of conserva- Conservation 124: 35–48. (5)
tion covenants. Conservation Letters. doi: 10.1111/ Hedrick, P. 2005. Large variance in reproductive success
conl.12193. (11) and the Ne/N ratio. Evolution 59: 1596–1599. (5)
Hare, M. P., L. Nunney, M. K., Schwartz, D. E. Ruzzante, Heleno, R. H., R. S. Ceia, J. A. Ramos, and J. Memmott.
et al. 2011. Understanding and estimating effective 2009. The effect of alien plants on insect abundance
population size for practical application in marine and biomass: a food web approach. Conservation
species management. Conservation Biology 25: Biology 23(2): 410–419. (4)
438–449. (5) Helfield, J. M., S. J. Capon, C. Nilsson, R. Jansson, and
Harrington, L. A., A. Moehrenschlager, M. Gelling, R. P. D. Palm. 2007. Restoration of rivers used for timber
D. Atkinson, et al. 2013. Conflicting and comple- floating: Effects on riparian plant diversity. Ecologi-
mentary ethics to animal welfare considerations in cal Applications 17: 840–851. (10)
reintroductions. Conservation Biology 27: 486–500. (7) Helm, D. 2015. Natural Capital: Valuing the Planet. Yale
Harris, J. A., R. J. Hobbs, E. Higgs, and J. Aronson. 2006. University Press, New Haven, CT. (3)
Ecological restoration and global climate change. Hendricks, S. A., P. R. S. Clee, R. J. Harrigan, J. P. Poll-
Restoration Ecology 14(2): 170–176. (10) inger, et al. 2016. Re-defining historical geographic
Harvell, D., R. Aronson, N. Baron, J. Connell, et al. 2004. range in species with sparse records: Implications
The rising tide of ocean diseases: Unsolved prob- for the Mexican wolf reintroduction program. Bio-
lems and research priorities. Frontiers in Ecology and logical Conservation 194, 48–57. (7)
the Environment 2: 375–382. (4) Heschel, M. S. and K. N. Paige. 1995. Inbreeding depres-
Hassett, B, M. Palmer, E. Bernhardt, S. Smith, et al. 2005. sion, environmental stress and population size
Restoring watersheds project by project: trends in variation in scarlet gilia (Ipomopsis aggregata). Con-
Chesapeake Bay tributary restoration. Frontiers in servation Biology 9: 126–133. (5)
Ecology and Environment 3: 259–267. (10) Hickey, V. and S. L. Pimm. 2011. How the World Bank
Hautier, Y., D. Tilman, F. Isbell, E. W. Seabloom, et al. funds protected areas. Conservation Letters 4:
2015. Anthropogenic environmental changes affect 269–277. (11)
ecosystem stability via biodiversity. Science 348: Higgs, E. 2003. Nature by Design: People, Natural Process,
336–340. (1, 3) and Ecological Restoration. MIT Press, Cambridge,
Havens, K., C. L. Jolls, J. E. Marik, P. Vitt, et al. 2012. MA. (10)
Effects of a non-native biocontrol weevil, Larinus
Bibliography 439

Higgs E., D. A. Falk, A. Guerrini, M. Hall, et al. 2014. The induced decline of an apex predator, the Tasmanian
role of history in restoration ecology. Frontiers in devil. Conservation Biology 28: 63–75. (2)
Ecology and the Environment 12(9): 499–506. (10) Hooke, R. L., J. F. Martín-Duque, and J. Pedraza. 2012.
Himes Boor, G. K. 2014. A framework for developing Land transformation by humans: a review. GSA
objective and measurable recovery criteria for Today 22(12): 4–10. (4)
threatened and endangered species. Conservation Horton, C., T. R. Peterson, P. Banerjee, and M. J. Peterson.
Biology 28: 33–43. (6) 2015. Credibility and advocacy in conservation sci-
Hinsley, A., D. Verissimo, and D. L. Roberts. 2015. ence. Conservation Biology 30(1), 23–32. doi: 10.1111/
Heterogeneity in consumer preferences for orchids cobi.12558. (1, 12)
in international trade and the potential for the use Horwitz, P. and C. M. Finlayson. 2011. Wetlands as set-
of market research methods to study demand for tings for human health: incorporating ecosystem
wildlife. Biological Conservation 190: 80–86. (4) services and health impact assessment into water
Hinz, H., V. Prieto, and M. J. Kaiser. 2009. Trawl distur- resource management. BioScience 61(9): 678–688. (3)
bance on benthic communities: Chronic effects and Huffington Post. 2015. Germany to Turn 62 Military
experimental predictions. Ecological Applications 19: Bases into Nature Sanctuary for Birds, Beetles and
761–773. (4) Bats. www.huffingtonpost.com/2015/06/19/german-
Hinz, H. L., M. Schwarzländer, A. Gassmann, and R. S. military-bases-nature-reserves_n_7623882.html (9)
Bourchier. 2014. Successes we may not have had: a Hughes, A. R., S. L. Williams, C. M. Duarte, K. L. Heck,
retrospective analysis of selected weed biological Jr., and M. Waycott. 2009. Associations of concern:
control agents in the United States. Invasive Plant Declining seagrasses and threatened dependent
Science and Management 7(4): 565–579. (10) species. Frontiers in Ecology and the Environment 7:
Hitzhusen, G. E. and M. E. Tucker. 2013. The potential of 242–246. (2)
religion for Earth Stewardship. Frontiers in Ecology Hughes, J. B. and J. Roughgarden. 2000. Species diversity
and the Environment 11: 368–376. (1) and biomass stability. American Naturalist 155:
Hobbs, R. J. and J. A. Harris. 2001. Restoration ecology: 618–627. (5)
repairing the earth’s ecosystems in the new millen- Hughes, T. P., H. U. I. Huang, and M. A. Young. 2013.
nium. Restoration Ecology 9(2): 239–246. (10) The wicked problem of China’s disappearing coral
Hobbs, R. J., D. N. Cole, L. Yung, E. S. Zavaleta, et al. reefs. Conservation Biology 27(2): 261–269. (4)
2010. Guiding concepts for park and wilderness Hughes, T. W. and K. Lee. 2015. The role of recreational
stewardship in an era of global environmental hunting in the recovery and conservation of the
change. Frontiers in Ecology and the Environment 8: wild turkey (Meleagris gallopavo spp.) in North
483–490. (8, 10) America. International Journal of Environmental Stud-
Hobbs, R. J., E. S. Higgs, and C. M. Hall. 2013. Novel ies. 72(5): 797–809. (7)
Ecosystems: Intervening in the New Ecological World Huhta, E. and J. Jokimäki. 2015. Landscape matrix frag-
Order. Wiley-Blackwell, Oxford, UK. (4, 10) mentation effect on virgin forest and managed for-
Hoeinghaus, D. J., A. A. Agostinho, L. C. Gomes, F. M. est birds: a multi-scale study. In J. A. Daniels (ed.),
Pelicice, et al. 2009. Effects of river impoundment on Advances in Environmental Research, Volume 36, pp.
ecosystem services of large tropical rivers: Embod- 95–111. Nova Science Publishers, Inc., Hauppauge,
ied energy and market value of artisanal fisheries. NY. (4)
Conservation Biology 23: 1222–1231. (3) Hulme, P. E. 2015. Invasion pathways at a crossroad:
Hoffman, S. W. and J. P. Smith. 2003. Population trends policy and research challenges for managing alien
of migratory raptors in western North America, species introductions. Journal of Applied Ecology
1977–2001. Condor 105: 397–419. (4) 52(6): 1418–1424. (4)
Hoffmann, M. T. Brooks, G. A. B. Fonseca, and J. M. C. Hultine, K. R., D. W. Bean, T. L. Dudley, and C. A. Geh-
Silva. 2008. Conservation planning and the IUCN ring. 2015. Species introductions and their cascad-
Red List. Endangered Species Research 6(2): 113–125. ing impacts on biotic interactions in desert riparian
(6) ecosystems. Integrative and Comparative Biology 55(4):
Hoffmann, M., C. Hilton-Taylor, A. Angulo, M. Böhm, et 587–601. (10)
al. 2010. The impact of conservation on the status of Human Microbiome Consortium. 2012. www.human-
the world’s vertebrates. Science 330: 1503–1509. (5) microbiome.org/ (2)
Holden, E. and K. G. Hoyer. 2005. The ecological foot- Humavindu, M. N. and J. Stage. 2015. Continuous finan-
print of fuels. Transportation Research, Part D 10: cial support will be needed. Animal Conservation
395–403. (4) 18(1): 18–19. (9)
Hole, D. G., B. Huntley, J. Arinaitwe, S. H. M. Butchart, Humphries, P. and K. O. Winemiller. 2009. Historical
et al. 2011. Toward a management framework for impacts on river fauna, shifting baselines, and chal-
networks of protected areas in the face of climate lenges for restoration. BioScience 59: 673–684. (10)
change. Conservation Biology 25: 305–315. (4) Hunter, M. O., M. Keller, D. Morton, B. Cook, et al. 2015.
Hollings, T., M. Jones, N. Mooney, and H. McCallum. Structural dynamics of tropical moist forest gaps.
2014. Trophic cascades following the disease- PLoS ONE 10(7): e0132144. (8)
440 Bibliography

Hylander, K., S. Nemomissa, J. Delrue, and W. Enkosa. protected areas in Britain. Biological Conservation
2013. Effects of coffee management on deforestation 142: 1515–1522. (8)
rates and forest integrity. Conservation Biology 27: Jacob, J., E. Jovic, and M. B. Brinkerhoff. 2009. Personal
1011–1019. (9) and planetary well-being: Mindfulness meditation,
pro-environmental behavior and personal quality of
Ikh Nart Nature Reserve. IkhNart.com (12) life in a survey from the social justice and ecological
Iknayan, K. J., M. W. Tingley, B. J. Furnas, S. R. Beiss- sustainability movement. Social Indicators Research
inger. 2014. Detecting diversity: emerging methods 93: 275–294. (3)
to estimate species diversity. Trends in Ecology and Jacobson, S. K. 2009. Communication Skills for Conserva-
Evolution. 29(2): 97–106. (2) tion Professionals, 2nd Ed. Island Press, Washington,
Indigenous Peoples Literature. 2009. indigenouspeople. DC. (12)
net (9) Jacobson, S. K., M. D. Mcduff, and M. Monroe. 2007. Pro-
Inouye, D. W., M. A. Morales, and G. J. Dodge. 2002. moting conservation through the arts: outreach for
Variation in timing and abundance of flowering by hearts and minds. Conservation Biology 21(1): 7–10. (1)
Delphinium barbeyi Huth (Ranunculaceae): the roles Jacobson, S. K., D. Wald, N. Haynes, and R. Sakurai.
of snowpack, frost, and La Nina, in the context of 2014. Wildlife communication and negotiation. In
climate change. Oecologia 130 (4): 543–550. (4) R. McCleery, C. Moorman and N. Peterson (eds.),
Intergovernmental Panel on Climate Change (IPCC). Urban Wildlife Science: Theory and Practice, Springer-
2013. Climate Change 2013: The Physical Science Basis. Verlag, New York. (12)
Contribution of Working Group I to the Fifth Assess- Jacquemyn, H., C. Van Mechelen, R. Brys, and O. Hon-
ment Report of the Intergovernmental Panel on Climate nay. 2011. Management effects on the vegetation
Change. T. F. Stocker and 9 others (eds.). Cambridge and soil seed bank of calcareous grasslands: An
University Press, Cambridge, UK. (4) 11-year experiment. Biological Conservation 144:
International Joint Commission. 2014. A Balanced Diet for 416–422. (8)
Lake Erie: Reducing Phosphorus Loadings and Harmful Jaeger, I., H. Hop, and G. W. Gabrielsen. 2009. Biomag-
Algal Blooms. Report of the Lake Erie Ecosystem nification of mercury in selected species from an
Priority. (10) Arctic marine food web in Svalbard. Science of the
International Plant Protection Convention (IPPC). 2014. Total Environment 407: 4744–4751. (4)
R. K. Pachauri and L. A. Meyer (eds.), Climate Jähnig, S. C., A. W. Lorenz, D. Hering, C. Antons, et al.
Change 2014: Synthesis Report. Contribution of 2011. River restoration success: A question of per-
Working Groups I, II and III to the Fifth Assessment ception. Ecological Applications 21: 2007–2015. (8)
Report of the Intergovernmental Panel on Climate Jäkäläniemi, A., A. H. Postila, and J. Tuomi. 2013. Accu-
Change. IPCC, Geneva. (4) racy of short-term demographic data in projecting
International Union for Conservation of Nature (IUCN) long-term fate of populations. Conservation Biology
2015. The IUCN Red List of Threatened Species. Ver- 27: 552–559. Many years of data are needed to build
sion 2015–4. www.iucnredlist.org Downloaded on 19 a good PVA model. (6)
November 2015. (1) Jamieson, I. G. 2011. Founder effects, inbreeding, and
International Union for Conservation of Nature (IUCN). loss of genetic diversity in four avian reintroduction
www.iucn.org (6, 8, 12) programs. Conservation Biology 25: 115–123. (5)
International Whaling Commission (IWC). www.iwc.int. (6) Janzen, D. H. 2001. Latent extinctions—the living dead.
International Work Group for Indigenous Affairs (IW- In S. A. Levin (ed.), Encyclopedia of Biodiversity, pp.
GIA). www.iwgia.org (9) 689–700. Academic Press, San Diego, CA. (5)
International Union for Conservation of Nature (IUCN). Jarošík, V., M. Konvička, P. Pyšek, T. Kadlec, and J. Beneš.
2001. IUCN Red List Categories and Criteria: 2011. Conservation in a city: Do the same principles
Version 3.1. IUCN Species Survival Commission. apply to different taxa? Biological Conservation 144:
IUCN, Gland, Switzerland. www.iucnredlist.org/ 490–499. (8)
technical-documents/categories-and-criteria/2001–cate- Jarvis, D. I., T. Hodgkin, A. H. Brown, J. Tuxill, et al. 2016.
gories-criteria (6) Crop Genetic Diversity in the Field and on the Farm:
International Union for Conservation of Nature (IUCN). Principles and Applications in Research Practices. Yale
2004. 2004 IUCN Red List of Threatened Species. University Press. (9)
www.iucnredlist.org (4) Jelinski, D. E. 2015. On a landscape ecology of a harle-
International Union for Conservation of Nature (IUCN). quin environment: the marine landscape. Landscape
2015. 2014 Annual Report of the Species Survival Com- Ecology 30(1): 1–6. (8)
mission and the Global Species Programme. portals.iucn. Jenkins, C. N. and L. Joppa. 2009. Expansion of the global
org/library/node/45591 (1) terrestrial protected area system. Biological Conserva-
tion 142: 2166–2174. (8)
Jackson, S. F., K. Walker, and K. J. Gaston. 2009. Relation- Jewell, A. 2013. Effect of monitoring technique on quality
ship between distributions of threatened plants and of conservation science. Conservation Biology 27:
501–508. (6)
Bibliography 441

Jones, C. G., J. H. Lawton, and M. Shachak. 1996. Organ- Committee on Government Reform, U. S. House of
isms as ecosystem engineers. In Ecosystem Man- Representatives, Washington, DC. (4)
agement, F. B. Samson and F. L. Knopf (eds.), pp. Kautz, R., R. Kawula, T. Hoctor, J. Comiskey, et al. 2006.
130–147. Springer, New York. (2) How much is enough? Landscape-scale conserva-
Jones, H. L. and J. M. Diamond. 1976. Short-time-base tion for the Florida panther. Biological Conservation
studies of turnover in breeding birds of the Califor- 130: 118–133. (9)
nia Channel Islands. Condor 76: 526–549. (6) Keeton, W. S., C. E. Kraft, and D. R. Warren. 2007. Mature
Jones, K. E., N. G. Patel, M. A. Levy, A. Storeygard, et al. and old-growth riparian forests: Structure, dynam-
2008. Global trends in emerging infectious diseases. ics, and effects on Adirondack stream habitats.
Nature 451: 990–994. (4) Ecological Applications 17: 852–868. (8)
Jones, L., A. Garbutt, J. Hansom, S. Angus. 2013. Impacts Keller R. P. and D. M. Lodge. 2007. Species invasions from
of climate change on coastal habitats. Marine Climate commerce in live aquatic organisms – problems and
Change Impacts Partnership: Science Review 2013: possible solutions. BioScience 57: 428–436. (4)
167–179. (4) Kelm, D. H., K. R. Wiesner, and O. von Helversen. 2008.
Jones, T. A. 2003. The restoration gene pool concept: be- Effects of artificial roosts for frugivorous bats on
yond the native versus non-native debate. Restora- seed dispersal in a Neotropical forest pasture mo-
tion Ecology 11(3): 281–290. (10) saic. Conservation Biology 22: 733–741. (2)
Joppa, L. N., S. R. Loarie, and S. L. Pimm. 2008. On the Keenelyside, K., N. Dudley, S. Cairns, C. Hall, and S.
protection of “protected areas.” Proceedings of the Stolton. 2012. Ecological restoration for protected
National Academy of Sciences USA 105: 6673–6678. (8) areas: principles, guidelines and best practices (Vol. 18).
Joppa, L. N. and A. Pfaff. 2009. High and far: biases in the IUCN. (10)
location of protected areas. PLOS ONE 4: e8273. (8) Kemp, D. R., H. Guodong, H. Xiangyang, D. L. Michalk,
Joppa, L. N., D. L. Roberts, and S. L. Pimm. 2011. The et al. 2013. Innovative grassland management
population ecology and social behavior of taxono- systems for environmental and livelihood benefits.
mists. Trends in Ecology and Evolution 26: 551–553. Proceedings of the National Academy of Sciences USA
(2, 6) 110: 8369–8374. (9)
Joppa, L. N., P. Visconti, C. N. Jenkins, and S. L. Pimm. Keppel, G., A. Naikatini, I. A. Rounds, R. L. Pressey, and
2013. Achieving the Convention on Biological N. T. Thomas. 2015. Local and expert knowledge
Diversity’s goals for plant conservation. Science 341: improve conservation assessment of rare and iconic
1100–1103. (8) Fijian tree species. Pacific Conservation Biology 21(3):
214–219. (6)
Jorge, M. L. S., M. Galetti, M. C. Ribeiro, and K. M. P.
Ferraz. 2013. Mammal defaunation as surrogate of Kiesecker, J. M., H. Copeland, A. Pocewicz, N. Nibbelink,
trophic cascades in a biodiversity hotspot. Biological et al. 2009. A framework for implementing biodi-
Conservation 163: 49–57. (2) versity offsets: selecting sites and determining scale.
BioScience 59: 77–84. (11)
Journey North. Annenberg Learner. www.learner.org/
jnorth (12) King, A. W., D. J. Hayes, D. N. Huntzinger, T. O. West,
et al. 2012. North American carbon dioxide sources
Junk, W. J., M. T. F. Piedade, L. Lourival, F. Wittmann, et
and sinks: magnitude, attribution, and uncertainty.
al. 2014. Brazilian wetlands: their definition, delin-
Frontiers in Ecology and the Environment 10: 512–519.
eation, and classification for research, sustainable
(3)
management, and protection. Aquatic Conservation
24(1): 5–22. (3) King, D. I., R. B. Chandler, J. M. Collins, W. R. Petersen,
et al. 2009. Effects of width, edge and habitat on the
abundance and nesting success of scrub-shrub birds
Kadoya, T., S. Suda, and I. Washitani. 2009. Dragonfly in powerline corridors. Biological Conservation 142:
crisis in Japan: A likely consequence of recent agri- 2672–2680. (9)
cultural habitat degradation. Biological Conservation
King, T., C. Chamberlan, and A. Courage. 2014. Assess-
142: 1899–1905. (8)
ing reintroduction success in long-lived primates
Kahler, J. S., G. J. Roloff, and M. L. Gore. 2013. Poaching through population viability analysis: western low-
risks in community-based natural resource manage- land gorillas Gorilla gorilla gorilla in Central Africa.
ment. Conservation Biology 27: 177–186. (22) Oryx 48(02): 294–303. (6)
Karanth, K. and R. DeFries. 2011. Nature-based tourism Kinnaird, M. F. and T. G. O’Brien. 2013. Effects of private
in Indian protected areas: New challenges for park land use, livestock management, and human toler-
management. Conservation Letters 4: 137–149. (3) ance on diversity, distribution, and abundance of
Karesh, W. B., R. A. Cook, E. L. Bennett, and J. Newcomb. large African mammals. Conservation Biology 27:
2005. Wildlife trade and global disease emergence. 1026–1039. (9)
CDC Emerging Infectious Diseases 11: 1000–1002. (4) Kintisch, K. 2013. Climate study highlights wedge issue.
Karl, T. R. 2006. Written statement for an oversight Science 339: 128–129. (4)
hearing: Introduction to Climate Change before the Kloor, K. 2015. The Battle for the Soul of Conservation
Science. Issues in Science and Technology 31: 74. (1)
442 Bibliography

Knapp, R. A., C. P. Hawkins, J. Ladau, and J. G. McClory. Archaea: Sequencing a Myriad of Type Strains.
2005. Fauna of Yosemite National Park lakes has PLoS Biology 12(8): e1001920. doi: 10.1371/journal.
low resistance but high resilience to fish introduc- pbio.1001920. (2)
tions. Ecological Applications 15: 835–847. (2)
Knoot, T. G., M. Rickenbach, and K. Silbernagel. 2015. Lacher, T. E., L. Boitani, and G. A. B. da Fonseca. 2012.
Payments for ecosystem services: will a new hook The IUCN global assessments: partnerships, col-
net more active family forest owners. Journal of laboration and data sharing for biodiversity science
Forestry 113(2): 210–218 (11) and policy. Conservation Letters 5: 327–333. (6)
Kobori, H. 2009. Current trends in conservation education Laikre, L., F. W. Allendorf, L. C. Aroner, C. S. Baker, et al.
in Japan. Biological Conservation 142: 1950–1957. (10) 2010. Neglect of genetic diversity in implementation
Kobori, H. and R. Primack. 2003. Participatory conserva- of the Convention on Biological Diversity. Conserva-
tion approaches for Satoyama: The traditional forest tion Biology 24: 86–88. (2)
and agricultural landscape of Japan. Ambio 32: Laikre, L., M. Jansson, F. W. Allendorf, S. Jakobsson, N.
307–311. (8) Ryman. 2013. Hunting effects on favourable con-
Köhler, H. R. and Triebskorn, R. 2013. Wildlife ecotoxi- servation status of highly inbred Swedish wolves.
cology of pesticides: can we track effects to the Conservation Biology 27: 248–253. (5)
population level and beyond? Science 341(6147): Laikre, L., M. K. Schwartz, R. S. Waples, and N. Ryman.
759–765. (4) 2010. Compromising genetic diversity in the wild:
Kolby, J. E., K. M. Smith, S. D. Ramirez, F. Rabemanan- Unmonitored large-scale release of plants and ani-
jara, F. et al. 2015. Rapid response to evaluate the mals. Trends in Ecology and Evolution 25: 520–529. (4)
presence of amphibian chytrid fungus (Batrachochy- Land Trust Alliance. www.landtrustalliance.org (11)
trium dendrobatidis) and ranavirus in wild amphib- Langton, M., L. Palmer, and Z. Ma Rhea. 2014. Com-
ian populations in Madagascar. PLoS ONE 10(6): munity-oriented protected areas for indigenous
e0125330. (7) peoples and local communities. In S. Stevens (ed.),
Korngold, J. 2008. God in the Wilderness: Rediscovering the Indigenous Peoples, National Parks, and Protected
Spirituality of the Great Outdoors with the Adventure Areas: A New Paradigm Linking Conservation, Culture,
Rabbi. Doubleday, New York. (1) and Rights, pp. 84–107. (8)
Krausmann, F., K-H Erb, S. Gingrich, H. Haberi, et al. Lapeyre, R. 2015. Wildlife conservation without financial
2013. Global human appropriation of net primary viability? The potential for payments for dispersal
production doubled in the 20th century. Proceedings areas’ services in Namibia. Animal Conservation
of the National Academy of Sciences 110(25): 10324– 18(1): 14–15. (9)
10329. (4) Larson, C. 2016. Shell trade pushes giant clams to the
Krauze-Gryz, D., J. Gryz,and J. Goszczyński, J. 2012. brink. Science 351: 323–324. (4)
Predation by domestic cats in rural areas of central Latin American and Caribbean Network of Environmen-
Poland: an assessment based on two methods. Jour- tal Funds (RedLAC). www.redlac.org (11)
nal of Zoology 288(4), 260–266. (4)
Laurance, W. F. 2008a. Adopt a forest. Biotropica 40: 3–6. (10)
Kroll, G. 2015. An environmental history of roadkill: road
ecology and the making of the permeable highway. Laurance, W. F. 2008b. Theory meets reality: how habitat
Environmental History 20(1), 4–28. (4) fragmentation research has transcended island
biogeographic theory. Biological Conservation 141:
Kross, S. M., J. M. Tylianakis, and X. J. Nelson. 2012. 1731–1744. (5)
Effects of introducing threatened falcons into vine-
yards on abundance of passeriformes and bird dam- Laurance, W. F. 2013. Does research help to safeguard
age to grapes. Conservation Biology 26: 142–149. (3) protected areas? Trends in Ecology and Evolution 28:
261–266. (12)
Kubiszewski, I., R. Costanzaa, C. Francob, P. Lawn, et al.
2013. Beyond GDP: Measuring and achieving global Laurance, W. F., M. Goosem, and S. G. W. Laurance. 2009.
genuine progress. Ecological Economics 93: 57–68. (3) Impacts of roads and linear clearings on tropical for-
ests. Trends in Ecology and Evolution 24: 659–679. (4)
Kuebbing, S. E., L. Souza, and N. J. Sanders. 2014. Effects
of co-occurring non-native invasive plant species on Laurance, W. F., H. Koster, M. Grooten, A. B. Anderson,
old-field succession. Forest Ecology and Management et al. 2012. Making conservation research more
324: 196–204. (4) relevant for conservation practitioners. Biological
Conservation 153: 164–168. (1)
Kueffer, C. and C. N. K. Kaiser-Bunbury. 2014. Reconcil-
ing conflicting perspectives for biodiversity conser- Laurance, W. F., S. G. Laurance, and D. W. Hilbert. 2008.
vation in the Anthropocene. Frontiers in Ecology and Long-germ dynamics of a fragmented rainforest
the Environment 12: 131–137. (10) mammal assemblage. Conservation Biology 22(5):
1154–1164. (5)
Kulkarni, M. V., P. M. Groffman, and J. B. Yavitt. 2008.
Solving the global nitrogen problem: It’s a gas! Fron- Laurance, W. F., T. E. Lovejoy, H. L. Vasconcelos, E. M.
tiers in Ecology and the Environment 6: 199–206. (4) Bruna, et al. 2002. Ecosystem decay of Amazonian
forest fragments: A 22–year investigation. Conserva-
Kyrpides, N. C., P. Hugenholtz, J. A. Eisen, T. Woyke, tion Biology 16: 605–618. (4)
et al. 2014. Genomic Encyclopedia of Bacteria and
Bibliography 443

Lavauden L. 1927. Les forêts du Sahara. Revue des Eaux et implications for primate conservation in Korup
Forêts 7(65): 329–41. (4) National Park, Cameroon. Biological Conservation
Lawler, J. J., S. P. Campbell, A. D. Guerry, M. B. Kolozs- 144: 738–745. (4)
vary, et al. 2002. The scope and treatment of threats Lindsey, P. A., G. A. Balme, P. J. Funston, P. Henschel, and
in endangered species recovery plans. Ecological L. T. Hunter. 2016. Life after Cecil: channelling global
Applications 12: 663–667. (4) outrage into funding for conservation in Africa.
Lawrence, A. J., R. Afif, M. Ahmed, S. Khalifa, and T. Conservation Letters. DOI: 10.1111/conl.12224 (9)
Paget. 2010. Bioactivity as an options value of sea Lindsey, P. A., G. Blame, M. Becker, C. Begg, et al. 2013.
cucumbers in the Egyptian Red Sea. Conservation The bushmeat trade in African savannas: impacts,
Biology 24: 217–225. (3) drivers, and possible solutions. Biological Conserva-
Le Roux, D. S., K. Ikin, D. B. Lindenmayer, G. Bistricer, tion 160: 80–96. (4)
et al. 2015. Enriching small trees with artificial Lindsey, P. A., C. P. Havemann, R. M. Lines, A. E. Price,
nest boxes cannot mimic the value of large trees et al. 2013. Benefits of wildlife-based land uses on
for hollow-nesting birds. Restoration Ecology. doi: private lands in Namibia and limitations affecting
10.1111/rec.12303. (8) their development. Oryx 47(01): 41–53. (9)
Lee, J. S. H., S. Abood, J. Ghahoul, B. Barus, et al. 2014. Linnell, J. D., P. Kaczensky, U. Wotschikowsky, N. Lescu-
Environmental impacts of large-scale oil palm enter- reux, and L. Boitani. 2015. Framing the relation-
prises exceed that of smallholdings in Indonesia. ship between people and nature in the context of
Conservation Letters 7: 25–33. (4) European conservation. Conservation Biology 29(4):
Leidner, A. K. and M. C. Neel. 2011. Taxonomic and 978–985. (1)
geographic patterns of decline for threatened and Liu, H., H. Ren, Q. Liu, X. Wen, M. Maunder, and J. Gao.
endangered species in the United States. Conserva- 2015. Translocation of threatened plants as a conser-
tion Biology 25: 716–725. (5, 6) vation measure in China. Conservation Biology, 29(6):
Leopold, A. 1949. A Sand County Almanac and Sketches 1537–1551. (7)
Here and There. Oxford University Press, New York. Liu, P., L. Sun, J. Li, L. Wang, et al. 2015. Population
(1, 3) viability analysis of Gloydius shedaoensis from
Leopold, A. 1953. Round River: From the Journals of Aldo northeastern China: A contribution to the assess-
Leopold. Oxford University Press, New York. (3) ment of the conservation and management status of
an endangered species. Asian Herpetological Research
Leopold, A. C. 2004. Living with the land ethic. BioScience
1(1): 48–56. (6)
54: 149–154. (1)
Liu, S. H., K. Li, and D. F. Hu. 2016. The incidence and
Lerner, J., J. Mackey and F. Casey. 2007. What’s in Noah’s
species composition of Gasterophilus (Diptera,
wallet? Land conservation spending in the United
Gasterophilidae) causing equine myiasis in northern
States. BioScience 57: 419–423. (8)
Xinjiang, China. Veterinary Parasitology 217: 36–38. (7)
Levin, S. A. (ed.). 2001. Encyclopedia of Biodiversity. Aca-
Lloyd, P., T. E. Martin, R. L. Redmond, U. Langer, and M.
demic Press, San Diego, CA. (2)
M. Hart. 2005. Linking demographic effects of habi-
Liebhold, A. M., E. G. Brockerhoff, L. J. Garrett, J. L. tat fragmentation across landscapes to continental
Parke J. L., and K. O. Britton. 2012. Live plant source-sink dynamics. Ecological Applications 15:
imports: the major pathway for forest insect and 1504–1514. (4)
pathogen invasions of the US. Frontiers in Ecology
Lloyd, R. A., K. A. Lohse, and T. P. A. Ferre. 2013. Influ-
and the Environment 10: 135–143. (4)
ence of road reclamation techniques on forest
Lin, J., D. Pan, S. J. Davis, Q. Zhang, et al. 2014. Chi- ecosystem recovery. Frontiers in Ecology and the
na’s international trade and air pollution in the Environment. 11: 75–81. (10)
United States. Proceedings of the National Academy
Loke, L. H., R. J. Ladle, T. J. Bouma, and P. A. Todd. 2015.
of Science USA. 111(5) 1736–1741. doi: 10.1073/
Creating complex habitats for restoration and recon-
pnas.1312860111. (11)
ciliation. Ecological Engineering 77: 307–313. (9)
Lindenmayer, D. and M. Hunter. 2010. Some guiding
Loss, S. R. and R. B. Blair. 2011. Reduced density and nest
concepts for conservation biology. Conservation Biol-
survival of ground-nesting songbires relative to
ogy 24: 1459–1468. (1)
earthworm invasions in northern hardwood forests.
Lindenmayer, D. B., G. E. Likens, A. Haywood, and L. Conservation Biology 25: 983–993. (4)
Miezis. 2011. Adaptive monitoring in the real world:
Lotze, H. K., M. Coll, A. M. Magera, C. Ward-Paige, and
Proof of concept. Trends in Ecology and Evolution 26:
L. Airoldi. 2011. Recovery of marine animal popula-
641–646. (8)
tions and ecosystems. Trends in Ecology and Evolution
Lindenmayer, D. B., A. Welsh, C. Donnelly, M. Crane, et 26: 595–605. (4)
al. 2009. Are nest boxes a viable alternative source of
Loucks, C., M. B. Mascia, A. Maxwell, K. Huy, et al. 2009.
cavities for hollow-dependant animals? Long-term
Wildlife decline in Cambodia, 1953–2005: Exploring
monitoring of nest box occupancy, pest use and at-
the legacy of armed conflict. Conservation Letters 2:
trition. Biological Conservation 142: 33–42. (8)
82–92. (4)
Linder, J. M. and J. F. Oates. 2011. Differential impact
Louda, S. M., D. Kendall, J. Connor, and D. Simberloff.
of bushmeat hunting on monkey species and
1997. Ecological effects of an insect introduced
444 Bibliography

for the biological control of weeds. Science 277: temperate rainforests. Conservation Biology 26:
1088–1090. (4) 238–247. (8)
Louisiana Coastal Wetlands Conservation and Restora- Magurran, A. E. 2013. Measuring Biological Diversity:
tion Task Force and the Wetlands Conservation and Frontiers in Measurement and Assessment. Oxford
Restoration Authority. 1998. Coast 2050: Toward a University Press, Oxford, UK. (2)
Sustainable Coastal Louisiana. Louisiana Department Malpai Borderlands Group. 2010. www.malpaiborderland-
of Natural Resources, Baton Rouge, LA. (10) sgroup.org (9)
Louv, R. 2005. Last Child in the Woods: Saving Our Children Máñez, K. S., G. Krause, I. Ring, and M. Glaser. 2014. The
from Nature-Deficit Disorder. Algonquin Books, Gordian knot of mangrove conservation: Disentan-
Chapel Hill, NC. (3) gling the role of scale, services and benefits. Global
Lovich, J. E., C. B. Yackulic, J. Freilich, M. Agha, et al. Environmental Change 28: 120–128. (4)
2014. Climatic variation and tortoise survival: Has a Manfredo, M. J., T. L. Teel, and A. M. Dietsch. 2016. Im-
desert species met its match? Biological Conservation plications of human value shift and persistence for
169: 214–224. (4) biodiversity conservation. Conservation Biology 30(2):
Loyd, K. A. T., S. M. Hernandez, J. P. Carroll, K. J. 287–296. (12)
Abernathy, and G. J. Marshall. 2013. Quantifying Mangel, M. and C. Tier. 1994. Four facts every conserva-
free-roaming domestic cat predation using animal- tion biologist should know about persistence. Ecol-
borne video cameras. Biological Conservation,160: ogy 75: 607–614. (5)
183–189. (4)
Marbuah, G., I. M. Gren, and B. McKie. 2014. Economics
Lu, Y., K. Wu, Y. Jiang, Y. Guo, et al. 2012. Widespread of harmful invasive species: a review. Diversity 6:
adoption of Bt cotton and insecticide decrease pro- 500–523. (4)
motes biocontrol services. Nature 487: 362–365. (4)
Maron, M., R. J. Hobbs, A. Moilanen, J. W. Matthews, et
Luck, G. W., P. Davidson, D. Boxall, and L. Smallbone. al. 2012. Faustian bargains? Restoration realities in
2011. Relations between urban bird and plant com- the context of biodiversity offset policies. Biological
munities and human well-being and connection to Conservation 158: 141–148. (10)
nature. Conservation Biology 25: 816–826. (3)
Maron, M., J. R. Rhodes, and P. Gibbons. 2013. Calculat-
Lunt, I. D., M. Byrne, J. J. Hellmann, N. J. Mitchell, et al. ing the benefit of conservation actions. Conservation
2013. Using assisted colonisation to conserve biodi- Letters 6: 359–367. (1, 3, 8, 12)
versity and restore ecosystem function under climate
Martin, T. G., S. Nally, A. A. Burbidge, S. Arnall, et al.
change. Biological Conservation 157: 172–177. (7)
2012. Acting fast helps avoid extinction. Conserva-
tion Letters 5: 274–280. (1, 5)
MacArthur, R. H. and E. O. Wilson. 1967. The Theory of Mascaro, J., R. F. Hughes, R. F., and S. A. Schnitzer. 2012.
Island Biogeography. Princeton University Press, Novel forests maintain ecosystem processes after
Princeton, NJ. (5, 8) the decline of native tree species. Ecological Mono-
Mace, G., H. Masundire, J. Baillie, T. Ricketts, et al. 2005. graphs 82(2): 221–228. (10)
Biodiversity. In R. Hassan, R. Scholes, and N. Ash Mascia, M. B. and C. A. Claus. 2009. A property rights
(eds.), Ecosystems and human well-being: Current state approach to understanding human displacement
and trends: Findings of the Condition and Trends Work- from protected areas: The case of Marine Protected
ing Group, pp. 77–122. Island Press, Washington, Areas. Conservation Biology 23: 16–23. (8)
DC. (5)
Mascia, M. B. and S. Pallier. 2011. Protected areas down-
Mack, R. N., D. Simberloff, W. M. Lonsdale, H. Evans, grading, downsizing, and degazettement (PADD)
et al. 2000. Biotic invasions: causes, epidemiology, and its conservation implications. Conservation
global consequences, and control. Ecological Applica- Letters 4: 9–20. (8)
tions 10(3): 689–710. (4)
Mascia, M. B., S. Pailler, M. L. Thieme, A., M. C. Bottrill,
MacKay, A., M. Allard, and M. A. Villard. 2014. Capac- et al. 2014. Commonalities and complementarities
ity of older plantations to host bird assemblages of among approaches to conservation monitoring and
naturally-regenerated conifer forests: a test at stand evaluation. Biological Conservation 169: 258–267. (11)
and landscape levels. Biological Conservation 170:
Mascia, M. B., S. Pallier, R. Krithivasan, V. Roshchanka,
110–119. (9)
et al. 2014. Protected area downgrading, downsiz-
Magera, A. M., J. E. M. Flemming, K. Kaschner, L. B. ing, and degazettement (PADDD) in Africa, Asia,
Christensen, H. K. Lotze. 2013. Recovery trends and Latin America and the Caribbean, 1900–2010.
in marine mammal populations. PLoS ONE doi: Biological Conservation 169: 355–361. (8)
10.1371/journal.pone.0077908. (4)
Matteson, K. C., D. J. Taron, and E. S. Minor. 2012. As-
Magnuson, J. J. 1990. Long-term ecological research and sessing citizen contributions to butterfly monitoring
the invisible present. BioScience 40: 495–501. (6) in two large cities. Conservation Biology 26: 557–564.
Magrach, A., A. R. Larrinaga, and L. Santamaría. 2012. (6)
Effects of matrix characteristics and interpatch Matthews, J., G. Van der Velde, A. B. De Vaate, F. P.
distance on functional connectivity in fragmented Collas, et al. 2014. Rapid range expansion of the
Bibliography 445

invasive quagga mussel in relation to zebra mussel McKinley, D. C., M. G. Ryan, R. A. Birdsey, C. P. Giar-
presence in The Netherlands and Western Europe. dina, et al. 2011. A synthesis of current knowledge
Biological invasions 16(1): 23–42. (4) on forests and carbon storage in the United States.
Matthews, T. J., F. Guilhaumon, K. A. Triantis, M. K. Ecological Applications 21: 1902–1924. (3)
Borregaard, and R. J. Whittaker. 2015. On the form McNeely, J. A. 1989. Protected areas and human ecology:
of species–area relationships in habitat islands and how national parks can contribute to sustaining
true islands. Global Ecology and Biogeography. doi: societies of the twenty-first century. In D. Western
10.1111/geb.12269. (8) and M. Pearl (eds.), Conservation for the Twenty-first
Maxted, N. 2001. Ex situ, in situ conservation. In S. A. Century, pp. 150–165. Oxford University Press, New
Levin (ed.), Encyclopedia of Biodiversity 2: 683–696. York. (9)
Academic Press, San Diego, CA. (7) McNeely, J. A., K. R. Miller, W. Reid, R. Mittermeier, et
Maynou, F. and J. E. Cartes. 2012. Effects of trawling on al. 1990. Conserving the World’s Biological Diversity.
fish and invertebrates from deep-sea coral facies of IUCN, World Resources Institute, CI, WWF-US, and
Isidella elongata in the western Mediterranean. Jour- the World Bank. Gland, Switzerland and Washing-
nal of the Marine Biological Association of the United ton, DC. (3)
Kingdom 92(7): 1501–1507. (4) McWilliams, J. 2013. Fertility treatments. Conservation 14:
McArthur, R. H. and E. O. Wilson. 1967. The Theory of Is- 34–39. (4)
land Biogeography. Princeton University Press, NJ. (5) Meffert, P. J. and F. Dziock. 2012. What determines occur-
McBride, M. F., S. T. Garnett, J. K. Szabo, A. H. Burbidge, rence of threatened bird species on urban waste-
et al. 2012. Structured elicitation of expert judg- lands? Biological Conservation 153: 87–100. (9)
ments for threatened species assessment: a case Meijaard, E., G., G. Albar, Nardiyono, Y. Rayadin, et al.
study on a continental scale using email. Methods in 2010. Unexpected ecological resilience in Bornean
Ecology and Evolution 3(5): 906–920. (6) orangutans and implications for pulp and paper
McCaffery, R., C. L. Richards-Zawacki, K. R. Lips. 2015. plantation management. PLoS ONE 5(9): e12813. (9)
The demography of Atelopus decline: Harlequin frog Meinhardt, K. A., and C. A. Gehring. 2013. Tamarix and
survival and abundance in central Panama prior to soil ecology. In A. Sher and M. Quigley (eds.), Tama-
and during a disease outbreak. Global Ecology and rix: A Case Study of Ecological Change in the American
Conservation 4: 232–242. (6) West. Oxford University Press New York. (4)
McCallum, M. L. 2015. Vertebrate biodiversity losses Meissen, J. C., S. M. Galatowitsch, and M. W. Cornett.
point to a sixth mass extinction. Biodiversity and Con- 2015. Risks of overharvesting seed from native tall-
servation 1–23. doi: 10.1007/s10531-015-0940-6. (5) grass prairies. Restoration Ecology 23(6): 882–891. (10)
McCarthy, D. P., P. F. Donald, J. P. W. Scharlemann, G. Melbourne, B. A. and A. Hastings. 2008. Extinction risk
M. Buchanan, et al. 2012. Financial costs of meeting depends strongly on factors contributing to stochas-
global biodiversity conservation targets: current ticity. Nature 454: 100–103. (5)
spending and unmet needs. Science 338(6109): Menges, E. S. 1992. Stochastic modeling of extinction
946–949. (11) in plant populations. In P. L. Fiedler and S. K. Jain
McCarthy, M. A., C. J. Thompson, and N. S. G. Williams. (eds.), Conservation Biology: The Theory and Practice
2006. Logic for designing nature reserves for mul- of Nature Conservation, Preservation and Management,
tiple species. American Naturalist 167: 717–727. (8) pp. 253–275. Chapman and Hall, New York. (5)
McCauley, D. J., M. L. Pinsky, S. R. Palumbi, J. A. Estes, Merckx, T. and H. M. Pereira. 2015. Reshaping agri-
et al. 2015. Marine defaunation: Animal loss in the environmental subsidies: From marginal farming to
global ocean. Science 347(6219): 1255641. (5) large-scale rewilding. Basic and Applied Ecology 16:
McCleery, R., J. A. Hostetler, and M. K. Oli. 2014. Better 95–103. (3)
off in the wild? Evaluating a captive breeding and Meyer, C. K., M. R. Whiles, and S. G. Baer. 2010. Plant
release program for the recovery of an endangered community recovery following restoration in
rodent. Biological Conservation 169: 198–205. (7) temporarily variable riparian wetlands. Restoration
McClelland, E. K. and K. A. Naish. 2007. What is the fit- Ecology 18: 52–64. (10)
ness outcome of crossing unrelated fish populations? Middleton, B. A. 2013. Rediscovering traditional vegeta-
A meta-analysis and an evaluation of future research tion management in preserves: trading experiences
directions. Conservation Genetics 8: 397–416. (5) between cultures and continents. Biological Conserva-
McClenachan, L., A. B. Cooper, K. E. Carpenter, and N. tion 158: 271–279. (8, 9)
K. Dulvy. 2012. Extinction risk and bottlenecks in Milder, J. C., J. P. Lassoie, and B. L. Bedford. 2008.
the conservation of charismatic marine species. Conserving biodiversity and ecosystem function
Conservation Letters 5: 73–80. (5) through limited development: An empirical evalua-
McConnachie, M. M., R. M. Cowling, B. W. van Wil- tion. Conservation Biology 22: 70–79. (11)
gen, and D. A. McConnachie. 2012. Evaluating the Millennium Ecosystem Assessment (MEA). 2005. Ecosys-
cost-effectiveness of invasive plant removal: a case tems and Human Well-being. 4 volumes. Island Press,
study from South Africa. Biological Conservation 155: Covelo, CA. (1, 2, 3, 4, 5, 9)
128–135. doi: 10.1016/j.biocon.2012.06.006. (4)
446 Bibliography

Miller, B., W. Conway, R. P. Reading, C. Wemmer, et al. under different climate change emission scenarios.
2004. Evaluating the conservation mission of zoos, Biological Conservation 145: 130–138. (6)
aquariums, botanical gardens, and natural history Molnar, J. L., R. L Gamboa, C. Revenga, and M. D. Spald-
museums. Conservation Biology 18: 86–93. (7) ing. 2008. Assessing the global threat of invasive
Miller, D. C., A. Agrawal, and J. T. Roberts. 2013. Biodi- species to marine biodiversity. Frontiers in Ecology
versity, governance, and the allocation of interna- and the Environment 9: 485–492. (4)
tional aid for conservation. Conservation Letters 6: Mora, C. 2009. Degradation of Caribbean coral reefs:
12–20. (11) focusing on proximal rather than ultimate drivers.
Miller, J. K., J. M. Scott, C. R. Miller, and L. P Waits. 2002. Reply to Rogers. Proceedings of the Royal Society of
The Endangered Species Act: Dollars and sense? London B: Biological Sciences 276(1655): 199–200. (4)
BioScience 52: 163–168. (6) Morandin, L. A. and C. Kremen. 2013. Hedgerow restora-
Miller, K. A., T. Bell, and J. M. Germano. 2014. Under- tion promotes pollinator populations and exports
standing publication bias in reintroduction biology native bees to adjacent fields. Ecological Applications
by assessing translocationsof New Zealand’s Her- 23: 829–839. (10)
petofauna. Conservation Biology. 28(4). doi: 10.1111/ Moreno-Mateos, D., V. Maris, A. Béchet, and M. Curran.
cobi.12254. (7) 2015a. The true loss caused by biodiversity offsets.
Miller, K. R. 1996. Balancing the Scales: Guidelines for Biological Conservation. 192: 552–559. (10)
Increasing Biodiversity’s Chances through Bioregional Moreno-Mateos, D., P. Meli, M. I. Vara-Rodríguez,
Management. World Resources Institute, Washing- and J. Aronson. 2015b. Ecosystem response to
ton, DC. (9) interventions: lessons from restored and created
Miller-Rushing, A. J. and R. B. Primack. 2008. Global wetland ecosystems. Journal of Applied Ecology 52(6):
warming and flowering times in Thoreau’s Concord: 1528–1537. (10)
A community perspective. Ecology 89: 332–341. (5) Moreno-Mateos, D., M. E. Power, F. A. Comín, and R.
Mills, S. 2003. Epicurean Simplicity. Island Press, Washing- Yockteng. 2012. Structural and functional loss in
ton, DC. (3) restored wetland ecosystems. PLoS-Biology 10(1):
Mills, M., J. G. Alvarez-Romero, K. Vance-Borland, P. 45. (4)
Cohen, et al. 2014. Linking regional planning and Morgan, J. L., S. E. Gergel, and N. C. Coops. 2010. Aerial
local action: towards using social network analy- photography: A rapidly evolving tool for ecological
sis in systematic conservation planning. Biological management. BioScience 60: 47–59. (8)
Conservation 169: 14–53. (8) Moro, M., A. Fischer, M. Czajkowski, D. Brennan, et al.
Mills Busa, J. H., 2013. Deforestation beyond borders: 2013. An investigation using the choice experiment
Addressing the disparity between production and method into options for reducing illegal bushmeat
consumption of global resources. Conservation Let- hunting in western Serengeti. Conservation Letters 6:
ters 6: 192–199. (4) 37–45. (4, 12)
Minteer, B. A. and J. P. Collins. 2008. From environmental Moro, D., M. W. Hayward, P. J. Seddon, and D. P. Arm-
to ecological ethics: toward a practical ethics for strong. 2015. Reintroduction biology of Australian
ecologists and conservationists. Science and Engineer- and New Zealand fauna: progress, emerging themes
ing Ethics 14: 483–501. (3) and future directions. Advances in Reintroduction
Minteer, B. A. and T. R. Miller. 2011. The new conserva- Biology of Australian and New Zealand Fauna 178. (7)
tion debate: Ethical foundations, strategic trade-offs, Morrell, V. 2014. Science behind plan to ease wolf protec-
and policy opportunities. Biological Conservation 144: tion is flawed, panel says. Science 343: 719. (7)
945–947. (11) Morrison, S. A. 2016. Designing virtuous socio-ecological
Minuzzi-Souza, T. T. C., N. Nitz, M. B. Knox, F. Reis, et cycles for biodiversity conservation. Biological Con-
al. 2016. Vector-borne transmission of Trypanosoma servation 195: 9–16. (12)
cruzi among captive Neotropical primates in a Bra- Morton, T. A. L., A. Thorn, J. M. Reed, R. Van Driesche, et
zilian zoo. Parasites and Vectors 9(1): 1. (7) al. 2015. Modeling the decline and potential recovery
Miththapala, S. 2008. Mangroves, Coastal Ecosystems Series. of a native butterfly following serial invasions by
Ecosystems and Livelihoods Group Asia, IUCN, exotic species. Biological Invasion 17:1683–1695. (5)
Colombo, Sri Lanka. (4) Moseley, L. (ed.). 2009. Holy Ground: A Gathering of Voices
Mitsch, W. J. and J. G. Gosselink. 2015. Wetlands. J. Wiley on Caring for Creation. Sierra Club Books, San Fran-
and Sons. New York. (4) cisco, CA. (3)
Mittermeier, R. A., P. R. Gil, M. Hoffman, J. Pilgrim, et al. Moss, A., E. Jensen, and M. Gusset. 2016. Probing the
2005. Hotspots Revisited: Earth’s Biologically Richest Link between biodiversity-related knowledge and
and Most Endangered Terrestrial Ecoregions. Conserva- self-reported pro-conservation behaviour in a global
tion International, Washington, DC. (6) survey of zoo visitors. Conservation Letters. doi:
Mlot, C. 2013. Are Isle Royale’s wolves chasing extinc- 10.1111/conl.12233. (7)
tion? Science 340: 919–921. (6) Mueller, J. G., I. H. B. Assanou, I. D. Guimbo, and A. M.
Molano-Flores, B. and T. J. Bell. 2012. Projected popula- Almedom. 2010. Evalutating rapid participatory
tion dynamics for a federally endangered plant rural appraisal as an assessment of ethnoecological
Bibliography 447

knowledge and local biodiversity patterns. Conser- Naess, A. 2008. The Ecology of Wisdom: Writings by Arne
vation Biology 24: 140–150. (6) Naess. A. Drengson and B. Devall (eds.). Counter-
Muir, J. 1901. Our National Parks. Houghton Mifflin, point, Berkeley, CA. (3)
Boston. (1) Naidoo, R., L. C. Weaver, R. W. Diggle, G. Matongo, et
Mukul, S. A. and J. Herbohn. 2016. The impacts of shift- al. 2016. Complementary benefits of tourism and
ing cultivation on secondary forests dynamics in hunting to communal conservancies in Namibia.
tropics: a synthesis of the key findings and spatio Conservation Biology. doi: 10.1111/cobi.12643. (9)
temporal distribution of research. Environmental Sci- Nakamura, K. and K. Takai. 2015. Indian Ocean Hydro-
ence and Policy 55: 167–177. (4) thermal Systems: Seafloor Hydrothermal Activities,
Munawar, S., M. F. Khokhar, and S. Atif. 2015. Reducing Physical and Chemical Characteristics of Hydrother-
emissions from deforestation and forest degradation mal Fluids, and Vent-Associated Biological Com-
implementation in northern Pakistan. International munities. In Ishibashi, J.-i. et al. (eds.), Subseafloor
Biodeterioration and Biodegradation 102: 316–323. (11) Biosphere Linked to Hydrothermal Systems. Springer,
Munilla, I., C. Diez, and A. Velando. 2007. Are edge bird Japan. doi: 10.1007/978-4-431-54865-2_12. (2)
populations doomed to extinction? A retrospective Namibia Association of CBNRM Support Organisations
analysis of the common guillemot collapse in Iberia. (NACSO). 2008. Namibia’s Communal Conservancies:
Biological Conservation 137: 359–371. (5) a review of progress in 2008. NACSO, Windhoek,
Munson, L., K. A. Terio, R. Kock, T. Mlengeya, et al. 2008. Namibia. (9)
Climate extremes promote fatal co-infections during Namibia Association of CBNRM Support Organisations
canine distemper epidemics in African lions. PLoS (NACSO). 2014. The state of community conserva-
ONE 3: e2545. (5) tion in Namibia—a review of communal conservancies,
Munson, S. M. and A. A. Sher. 2015. Long-term shifts community forests and other CBNRM initiatives (2013
in the phenology of rare and endemic Rocky Annual Report). Windhoek: Namibia Association of
Mountain plants. American Journal of Botany 102(8): CBNRM Support Organizations. doi: 10.1111/j.1523-
1268–1276. (4) 1739.2012.01960.x. (9)
Muradian, R., M. Arsel, L. Pellegrini, F. Adaman, et al. Naniwadekar, R., Shukla, U., Isvaran, K., Datta, A. 2015.
2014. Payments for ecosystem services and the fatal Reduced hornbill abundance associated with low
attraction of win-win solutions. Conservation Letters seed arrival and altered recruitment in a hunted and
6: 274–279. (11) logged tropical forest. PLoS ONE 10(3) e0120062.
doi: 10.1371/journal.pone.0120062. (2)
Murcia, C. 1995. Edge effects in fragmented forests:
implications for conservation. Trends in Ecology and Nash, S. 2009. Ecotourism and other invasions. BioScience
Evolution 10(2): 58–62. (4) 59: 106–110. (3)
Murcia, C., J. Aronson, G. H. Kattan, D., Moreno-Mateos, et National Geographic. 2015. Graphic Shows Who’s Buying
al. 2014. A critique of the “novel ecosystem” concept. and Selling Animals Globally. news.nationalgeo-
Trends in Ecology and Evolution 29(10): 548–553. (10) graphic.com/2015/06/150615-data-points-infographic-
animal-trade (4)
Murray-Smith, C., N. A. Brummitt, A. T. Oliveira-Filho,
S. Bachman, et al. 2009. Plant diversity hotspots in National Marine Sanctuary of American Samoa. ameri-
the Atlantic coastal forests of Brazil. Conservation cansamoa.noaa.gov (9)
Biology 23: 151–163. (8) National Oceanic and Atmospheric Administration
Musiani, M., C. Mamo, L. Boitani, C. Callaghan, et al. (NOAA). 2015. National Centers for Environ-
2003. Wolf depredation trends and the use of fladry mental Information, State of the Climate: Global
barriers to protect livestock in western North Analysis for August 2015. www.ncdc.noaa.gov/sotc/
America. Conservation Biology 17: 1538–1547. (7) global/201508 (4)
Myers, N., N. Golubiewski, and C. J. Cleveland. 2007. National Oceanic and Atmospheric Administration
Perverse subsidies. In C. J. Cleveland (ed.), Encyclo- (NOAA). 2016. National Centers for Environ-
pedia of Earth. Environmental Information Coalition, mental Information, State of the Climate: Global
National Council for Science and the Environment, Analysis for January 2016. www.ncdc.noaa.gov/sotc/
Washington, DC. www.eoearth.org/article/Perverse_ global/201601. (4)
subsidies (3) Natural Resource Stewardship and Science. 2015. Office
of the Associate Director. Strategic framework for
National Park Service Research Learning Centers.
Nabhan, G. P. 2008. Where Our Food Comes From: Retracing National Park Service, Natural Resource Steward-
Nicolay Vavilov’s Quest to End Famine. Island Press, ship and Science, Washington, DC. www.nature.nps.
Washington, DC. (7) gov/rlc /framework.pdf (12)
Naeem, S, A. V. Ingram, T. Agardy, P. Barten, et al. 2015. Native Seeds SEARCH: Southwestern Endangered
Get the science right when paying for nature’s Aridland Resource Clearing House. 2009. www.
services. Science 347(6227): 1206–1207. (11) nativeseeds.org (9)
Naess, A. 1989. Ecology, Community and Lifestyle. Cam- NatureServe Explorer. 2009. www.natureserve.org/explorer (6)
bridge University Press, Cambridge, MA. (3)
448 Bibliography

NatureServe: A Network Connecting Science with Con- avoided deforestation success of protected areas
servation. 2009. www.natureserve.org (6) in the Brazilian Amazon. Proceedings of the National
Nee, S. 2003. Unveiling prokaryotic diversity. Trends in Academy of Sciences USA 110: 4956–5961. (8, 11)
Ecology and Evolution 18: 62–63. (2) Noon, B. R., L. L. Bailey, T. D. Sisk, and K. S. McKelvey.
Negrón-Ortiz, N. 2014. Pattern of expenditures for plant 2012. Efficient species-level monitoring at the land-
conservation under the Endangered Species Act. scape scale. Conservation Biology 26: 432–441. (6)
Biological Conservation 171: 36–43. (11) Norris, K., A. Asase, B. Collen, J. Gockowski, et al. 2010.
Nelson, E. J., P. Kareiva, M. Ruckelshaus, K. Arkema, et. Biodiversity in a forest-agriculture mosaic—The
al. 2013. Climate change’s impact on key ecosystem changing face of West African rainforests. Biological
services and the human well-being they support in Conservation 143: 2341–2350. (9)
the US. Frontiers in Ecology and the Environment. 11: Nuñez, T. A., J. J. Lawler, B. H. Mcrae, D. J. Pierce, et
483–493. (4) al. 2013. Connectivity planning to address climate
Nelson, M. P., J. T. Bruskotter, J. A. Vucetich, and G. change. Conservation Biology 27: 407–416. (4, 8)
Chapron. 2016. Emotions and the Ethics of Conse- North American Breeding Bird Survey (BBS). www.pwrc.
quence in Conservation Decisions: Lessons from usgs.gov/bbs. (6)
Cecil the Lion. Conservation Letters. doi: 10.1111/
conl.12232 (9) Oakleaf, J. R., C. M. Kennedy, S. Baruch-Mordo, P. C.
Nepstad, D., S. Schwartzman, B. Bamberger, M. Santilli, West, et al. 2015. A World at Risk: Aggregating
et al. 2006. Inhibition of Amazon deforestation and Development Trends to Forecast Global Habitat
fire by parks and indigenous lands. Conservation Conversion. PLoS ONE 10(10): e0138334. (4)
Biology 20: 65–73. (4) Odell, J., M. E. Mather, and R. M. Muth. 2005. A biosocial
Newbold, S. C. and J. V. Siikamäki. 2009. Prioritizing approach for analyzing environmental conflicts: A
conservation activities using reserve site selection case study of horseshoe crab allocation. BioScience
methods and population viability analysis. Ecologi- 55: 735–748. (3)
cal Applications 19: 1774–1790. (3) O’Donnell, C. F. and J. M. Hoare. 2012. Quantifying the
Newmark, W. D. 1995. Extinction of mammal popula- benefits of long-term integrated pest control for
tions in western North American national parks. forest bird populations in a New Zealand temperate
Conservation Biology 9: 512–527. (8) rainforest. New Zealand Journal of Ecology 131–140.
Newmark, W. D. 2008. Isolation of African protected (4)
areas. Frontiers in Ecology and the Environment 6: Ogden, L. E. 2015. Do Wildlife Corridors Have a Down-
321–328. (8, 9) side? BioScience 65(4): 452–452. (8)
Nicoll, M. A. C., C. G. Jones, and K. Norris. 2004. Okin, G. S., A. Parsons, J. Wainwright, J. E. Herrick, et al.
Comparison of survival rates of captive-reared and 2009. Do changes in connectivity explain desertifica-
wild-bred Mauritius kestrels (Falco punctatus) in a tion? BioScience 59: 237–244. (4)
re-introduced population. Biological Conservation Oldekop, J. A., G. Holmes, W. E. Harris, and K. L. Evans.
118: 539–548. (7) 2015. A global assessment of the social and conser-
Nieto-Romero, M., A. Milcu, J. Leventon, F. Mikulcak, vation outcomes of protected areas. Conservation
and J. Fischer. 2016. The role of scenarios in foster- Biology 30(1): 133–141. (8)
ing collective action for sustainable development: Olden, J. D., M. J. Kennard, J. J. Lawler, and N. L. Poff.
Lessons from central Romania. Land Use Policy 50: 2011. Challenges and opportunities in implementing
156–168. (11) managed relocation for conservation of freshwater
Nijman, V., K. A. I. Nekaris, G. Donati, M. Bruford, and J. species. Conservation Biology 25: 40–47. (7)
Fa. 2011. Primate conservation: Measuring and miti- Oldfield, T. E. E., R. J. Smith, S. R. Harrop, and N. Lead-
gating trade in primates. Endangered Species Research er-Williams. 2003. Field sports and conservation in
13: 159–161. (4) the United Kingdom. Nature 423: 531–533. (3)
Niles, L. J. 2009. Effects of horseshoe crab harvest in Olivieri, I., J. Tonnabel, O. Ronce, and A. Mignot. 2016.
Delaware Bay on red knots: Are harvest restrictions Why evolution matters for species conservation: per-
working? BioScience 59: 153–164. (3) spectives from three case studies of plant metapopu-
Nimmo, D. G., L. T. Kelly, L. M. Spence-Bailey, S. J. Wat- lations. Evolutionary Applications 9(1): 196–211. (7)
son, et al. 2013. Fire mosaics and reptile conserva- O’Meilla, C. 2004. Current and reported historical range
tion in a fire-prone region. Conservation Biology 27: of the American burying beetle. U. S. Fish and Wild-
345–353. (8) life Services, Oklahoma Ecological Services Field
Noël, F., N. Machon, and A. Robert. 2013. Integrating Office, OK. (5)
demographic and genetic effects of connections O’Neill, B. C., B. Liddle, L. Jiang, K. R. Smith, K. R., et
on the viability of an endangered plant in a highly al. 2012. Demographic change and carbon dioxide
fragmented habitat. Biological Conservation 158: emissions. The Lancet 380(9837): 157–164. (4)
167–174. (6)
Oppel, S., B. M. Beaven, M. Bolton, J. Vickery, and T. W.
Nolte, C., A. Agrawal, K. M. Silvus, and B. S. Soares-Fil- Body. 2011. Eradication of invasive mammals on
ho. 2013. Governance regime and location influence
Bibliography 449

islands inhabited by humans and domestic animals. syndrome in conservation. Conservation Letters 2:
Conservation Biology 25: 232–240. (4) 93–100. (6)
Oppenheimer, J. D., S. K. Beaugh, J. A. Knudson, P. Muel- Pardini, R., S. M. de Souza, R. Braga-Neto, and J. P.
ler, et al. 2015. A collaborative model for large-scale Metzger. 2005. The role of forest structure, fragment
riparian restoration in the western United States. size and corridors in maintaining small mammal
Restoration Ecology 23(2): 143–148. (10) abundance and diversity in an Atlantic forest land-
Ore, J., S. Elbaum, A. Burgin, and C. Detweiler. 2015. scape. Biological Conservation 12: 253–266. (8)
Autonomous aerial water sampling. In L. Mejias, P. Parlato, E. H. and D. P. Armstrong. 2013. Predicting post-
Corke, and J. Roberts (eds.) Field and Service Robot- release establishment using data from multiple rein-
ics, pp 137–151. Springer International Publishing, troductions. Biological Conservation 160: 97–104. (7)
Switzerland. (6) Pawlowski, J., Audic, S., Adl, S., Bass, D., et al. 2012.
Orgnization for Economic Co-operation and Develop- CBOL protist working group: barcoding eukaryotic
ment. www.oecd.org (11) richness beyond the animal, plant, and fungal king-
Orrock, J. L. and E. I. Damschen. 2005. Corridors cause doms. PLoS Biology 10(11): e1001419. (2)
differential seed predation. Ecological Applications PBL Netherlands Environmental Assessment Agency.
15: 793–798. (8) 2012. Trends in Global CO2 Emissions 2012 Report.
Österblom, H. and Ö Bodin. 2012. Global cooperation PBL Netherlands Environmental Assessment Agen-
among diverse organizations to reduce illegal fish- cy The Hague The Netherlands, European Commis-
ing in the southern ocean. Conservation Biology 26: sion Joint Research Centre Institute for Environment
638–648. (4) and Sustainability. doi: 10.2788/33777. (4)
Osterlind, K. 2005. Concept formation in environmental Peace Parks Foundation. www.peaceparks.org (11)
education: 14–year olds’ work on the intensified Peakall, R., D. Ebert, L. J. Scott, P. F. Meagher, and C. A.
greenhouse. International Journal of Science Education Offord. 2003. Comparative genetic study confirms
27: 891–908. (3) exceptionally low genetic variation in the ancient
Ostfeld, R. S. 2009. Climate change and the distribu- and endangered relictual conifer, Wollemia nobilis
tion and intensity of infectious diseases. Ecology 4: (Araucariaceae). Molecular Ecology 12: 2331–2343. (5)
903–905. (4) Pe’er, G., M. A. Tsianou, K. W. Franz, Y. G. Matsinos, et
Ottewell, K., J. Dunlop, N. Thomas, K. Morris, et al. 2014. al. 2014. Toward better application of minimum area
Evaluating the success of translocations in main- requirements in conservation planning. Biological
taining genetic diversity in a threatened mammal. Conservation 170: 92–102. (5, 6)
Biological Conservation 171: 209–219. (7) Peery, M. Z., S. R. Beissinger, S. H. Newman, E. B.
Burkett, and T. D. Williams. 2004. Applying the
declining population paradigm: Diagnosing causes
Packer, C. 2015. Lions in the Balance: Man-Eaters, Manes,
of poor reproduction in the marbled murrelet. Con-
and Men with Guns. University of Chicago Press. (5)
servation Biology 18: 1088–1098. (5)
Packer, C., A. Loveridge, S. Canney, T. Caro, et al. 2013.
Peh, K. S. H., J. de Jong, N. S. Sodhi, S. L. H. Lim, et al.
Conserving large carnivores: Dollars and fence.
2005. Lowland rainforest avifauna and human
Conservation Letters 16: 635–641. (8)
disturbance: persistence of primary forest birds in
Paknia, O., H. Rajaei, and A. Koch. 2015. Lack of well- selectively logged forests and mixed-rural habitats
maintained natural history collections and taxono- of southern Peninsular Malaysia. Biological Conserva-
mists in megadiverse developing countries hampers tion 123: 489–505. (9)
global biodiversity exploration. Organisms Diversity
Peres, C. A. and M. Schneider. 2012. Subsidized agricul-
and Evolution 15(3): 619–629. (12)
tural resettlements as drivers of tropical deforesta-
Palomares, F., J. A. Godoy, J. V. Lόpez-Bao, A. Rodriguez, tion. Biological Conservation 151: 65–68. (4)
et al. 2012. Possible extinction vortex for a popula-
Perring, M. P., P. Audet, and D. Lamb. 2014. Novel eco-
tion of Iberian lynx on the verge of extirpation.
systems in ecological restoration and rehabilitation:
Conservation Biology 26: 689–697. (5)
Innovative planning or lowering the bar? Ecological
Pan, Y., R. A. Birdsey, J. Fang, R. Houghton, et al. 2011. Processes 3(1): 1–4. (10)
A large and persistent carbon sink in the world’s
Perry, G. and D. Vice. 2009. Forecasting the risk of brown
forests. Science 333: 988–993. (3)
tree snake dispersal from Guam: A mixed transport-
Pandolfi, J. M., S. R. Connoly, D. J. Marshall, and A. L. establishment model. Conservation Biology 23:
Cohen. 2011. Projecting coral reef futures under 992–1000. (4)
global warming and ocean acidification. Science 333:
Peterson, M. J., D. M. Hall, A. M. Feldpausch-Parker, and
418–442. (4)
T. R. Peterson. 2010. Obscuring ecosystem func-
Papadopoulou, A., D. Chesters, I. Coronado, G. De La tion with the application of the ecosystem services
Cadena, et al. 2015. Automated DNA-based plant concept. Conservation Biology 24: 113–119. (3)
identification for large-scale biodiversity assess-
Phelps, J., and E. L. Webb. 2015. “Invisible” wildlife
ment. Molecular Ecology Resources 15: 136–152. (2)
trades: Southeast Asia’s undocumented illegal trade
Papworth, S. K., J. Rist, L. Coad, and E. J. Milner-
Gulland. 2009. Evidence for shifting baseline
450 Bibliography

in wild ornamental plants. Biological Conservation disciplinarity in conservation science. Conservation


186: 296–305. (3) Biology 28: 22–32. (1)
Phelps, J., L. R. Carrasco, E. L. Webb, L. P. Koh, et al. Porensky, L. M. and T. P. Young. 2013. Edge effect
2013. Agricultural intensification escalates future interactions in fragmented and patchy landscapes.
conservation costs. Proceedings of the National Acad- Conservation Biology 27: 509–519. (4)
emy of Sciences USA 110: 7601–7606. (9, 11) Porszt, E. J., R. M. Peterman, N. K. Dulvy, A. B. Cooper,
Philpott, S. M., P. Bichier, R. Rice, and R. Greenberg. 2007. et al. 2012. Reliability of indicators of decline in
Field-testing ecological and economic benefits of abundance. Conservation Biology 26: 894–904. (6)
coffee certification programs. Conservation Biology Posey, D. A. and M. J. Balick (eds.). 2006. Human Impacts
21: 975–985. (9) on Amazonia: The Role of Traditional Ecological
Phua, M. H., S. Tsuyuki, N. Furuya, and J. S. Lee. 2008. Knowledge in Conservation and Development.
Detecting deforestation with a spectral change de- Columbia University Press, New York. (8, 9)
tection approach using multitemporal Landsat data: Possingham, H. P., M. Bode, M. and C. J. Klein, C. J. 2015.
A case study of Kinabalu Park, Sabah, Malaysia. Optimal conservation outcomes require both resto-
Journal of Environmental Management 88: 784–795. (4) ration and protection. PLoS Biology 13(1): e1002052–
Pikesley, S. K., B. J. Godley, H. Latham, P. B. Richardson, e1002052. (8)
et al. 2016. Pink sea fans (Eunicella verrucosa) as Poudel, R. C., M. Möller, J. Liu, L.-M. Gao, et al. 2014. Low
indicators of the spatial efficacy of Marine Protected genetic diversity and high inbreeding of the endan-
Areas in southwest UK coastal waters. Marine gered yews in Central Himalaya: implications for
Policy 64: 38–45. (9) conservation of their highly fragmented populations.
Pilgrim, J. D., S. Brownlie, J. M. M. Ekstrom, T. A. Gard- Diversity and Distributions 20(11): 1270–1284. (5)
ner, et al. 2013. A process for assessing the offsetabil- Powell, K. I., J. M. Chase, and T. M. Knight. 2013.
ity of biodiversity impacts. Conservation Letters 6: Invasive plants have scale-dependent effects on di-
376–384. (11, 12) versity by altering species-area relationships. Science
Pimentel, D., R. Zuniga, and D. Morrison. 2005. Update 339: 316–318. (3)
on the environmental and economic costs associated Power, M. E., D. Tilman, J. A. Estes, B. A. Menge, et al.
with alien-invasive species in the United States. 1996. Challenges in the quest for keystones. BioSci-
Ecological Economics 52: 273–288. (4) ence 46: 609–620. (2)
Pimm, S. L, C. N. Jenkins, R. Abell, T. M. Brooks, et al. Power, T. M. and R. N. Barret. 2001. Post-Cowboy Econom-
2014. The biodiversity of species and their rates of ics: Pay and Prosperity in the New American West.
extinction, distribution, and protection. Science 344: Island Press, Washington, DC. (3)
1246752. (1, 5)
Prescott-Allen, C. and R. Prescott-Allen. 1986. The First
Pimm, S. L. and C. Jenkins. 2005. Sustaining the variety Resource: Wild Species in the North American Economy.
of life. Scientific American 293(33): 66–73. (4, 5) Yale University Press, New Haven, CT. (3)
Pimm, S. L. and L. N. Joppa. 2015. How Many Plant Primack, R. B. and A. J. Miller-Rushing. 2012. Uncover-
Species are There, Where are They, and at What ing, collecting, and analyzing records to investigate
Rate are They Going Extinct? Annals of the Missouri the ecological impacts of climate change: A template
Botanical Garden 100(3): 170–176. (5) from Thoreau’s Concord. BioScience 62: 170–181. (4)
Pinchot, G. 1947. Breaking New Ground. Harcourt Brace, Primack, R. B., A. J. Miller-Rushing, and K. Dhara-
New York. (1) neeswaran. 2009. Changes in the flora of Thoreau’s
Pittman, S. E., M. S. Osbourn, and R. D. Semlitsch. 2014. Concord. Biological Conservation 142: 500–508. (5)
Movement ecology of amphibians: A missing Protected Planet. www.wdpa.org (8)
component for understanding population declines.
Purvis, A., J. L. Gittleman, G. Cowlishaw, and G. M.
Biological Conservation 169: 44–53. (6)
Mace. 2000. Predicting extinction risk in declining
Pocock, M. J., S. E. Newson, I. G. Henderson, J. Peyton, species. Proceedings of the Royal Society of London B:
et al. 2015. Developing and enhancing biodiversity Biological Sciences 267(1456): 1947–1952. (5)
monitoring programmes: a collaborative assess-
ment of priorities. Journal of Applied Ecology 52(3):
686–695. (8) Quammen, D. 1996. The Song of the Dodo: Island Biogeog-
raphy in an Age of Extinctions. Scribner, New York. (5)
Polidoro, B. A., K. E. Carpenter, L. Collins, N. C. Duke,
et al. 2010. The loss of species: Mangrove extinction Quayle, J. F., L. R. Ramsay, and D. F Fraser 2007. Trend
risk and geographic areas of global concern. PLoS in the status of breeding bird fauna in British Co-
ONE 5(4): e10095. (4) lumbia, Canada, based on the IUCN Red List Index
method. Conservation Biology 21(5): 1241–1247. (6)
Polishchuk, L. V., K. Y. Popadin, M. A. Baranova, and A.
S. Kondrashov. 2015. A genetic component of extinc- QUINTESSENCE Consortium. 2016. Networking our
tion risk in mammals. Oikos 124(8): 983–993. (5) way to better ecosystem service provision. Trends in
Ecology and Evolution 31(2): 105–115. (12)
Pooley, S. P., J. A. Mendelsohn, and E. J. Milner-Gulland.
2014. Hunting down the chimera of multiple
Bibliography 451

Radeloff, V. C., S. I. Stewart, T. J. Hawbaker, U. Gimmi, et die-offs in vertebrates. Animal Conservation 6(02):
al. 2010. Housing growth in and near United States 109–114. (5)
protected areas limits their conservation value. Reed, J. M. 1999. The role of behavior in recent avian
Proceedings of the National Academy of Sciences USA extinctions and endangerments. Conservation Biology
107: 940–945. (9) 13: 232–241. (5)
Raghavan, R., N. Dahanukar, M. F. Tlusty, A. L. Rhyne, Reed, J. M., C. S. Elphick, E. N. Ieno, and A. F. Zuur. 2011.
et al. 2013. Uncovering an obscure trade: threatened Long-term population trends of endangered Hawai-
freshwater fishes and the aquarium pet markets. ian waterbirds. Population Ecology 53: 473–481. (6)
Biological Conservation,164: 158–169. (4) Reed, J. M., C. S. Elphick, A. F. Zuur, E. N. Ieno, and G.
Rai, N. D. and K. S. Bawa. 2013. Insering politics and M. Smith. 2007. Time series analysis of Hawaiian
history in conservation. Conservation Biology 27: waterbirds. In A. F. Zuur, E. N. Ieno, and G. M.
425–428. (9) Smith (eds.), Analysis of Ecological Data, pp. 613–632.
Ramakrishnan, U., J. A. Santosh, U. Ramakrishnan, and Springer-Verlag, the Netherlands. (6)
R. Sukumar. 1998. The population and conserva- Reed, S. E., J. A. Hilty, and D. M. Theobald. 2014. Guide-
tion status of Asian elephants in the Periyar Tiger lines and incentives for conservation development
Reserve, southern India. Current Science India 74: in local land-use regulations. Conservation Biology
110–113. (5) 28: 258–268. (11)
Ramsar Convention Secretariat. www.ramsar.org (11) Reeve, R. 2014. Policing International Trade in Endangered
Rands, M. R., W. M. Adams, L. Bennun, S. H. M. Species: The CITES Treaty and Compliance. Routledge,
Butchart, et al. 2010. Biodiversity conservation: London. (6)
Challenges beyond 2010. Science 329: 1298–1303. (11) Régnier, C., B. Fontaine, and P. Bouchet. 2009. Not know-
Randolph, J. and G. M. Masters. 2008. Energy for Sustain- ing, not recording, not listing: Numerous unno-
ability: Technology, Planning, Policy. Island Press, ticed mollusk extinctions. Conservation Biology 23:
Washington, DC. (4) 1214–1221. (5, 6)
Raup, D. M. 1988. Diversity Crises in the Geological Past. Regos, A., M. D’Amen, N. Titeux, S. Herrando, A.
In E. O. Wilson and F. M Peter (eds.), Biodiversity, Guisan, and L. Brotons. 2016. Predicting the future
Chapter 5. National Academies Press, Washington, effectiveness of protected areas for bird conserva-
DC. (5) tion in Mediterranean ecosystems under climate
Raup, D. M. and Sepkoski, J. J. 1982. Mass extinctions in change and novel fire regime scenarios. Diversity
the marine fossil record. Science 215(4539): 1501– and Distributions, 22(1): 83–96. (8)
1503. (5) Regueira, R. F. S. and E. Bernard. 2012. Wildlife sinks.
Reading, R. 2015. Mongolia Field Program Update. Associa- Quantifying the impact of illegal bird trade in
tion of Zoos and Aquariums (AZA) 2015 Mid-year street markets in Brazil. Biological Conservation 149:
Meeting, Columbia, SC. (12) 16–22. (4)
Reading, R. R., T. J. Weaver, J. R. Garcia, R. Elias Piperis, Riehl, B., H. Zerriffi, and R. Naidoo. 2015. Effects of
et al. (eds.). 2011. Lake Titicaca’s Frog (Telmatobius community-based natural resource management on
culeus) Conservation Strategy Workshop. Decem- household welfare in Namibia. PLoS ONE 10(5). (9)
ber 13–15, 2011. Bioscience School, Universidad Reiker, J., B. Schulz, V. Wissemann, and B. Gemeinhol-
Nacional del Altiplano, Puno, Peru. Conservation zer. 2015. Does origin always matter? Evaluating
Breeding Specialist Group CBSG/(SSC/IUCN) the influence of nonlocal seed provenances for
Mesoamerica. (4) ecological restoration purposes in a widespread
Redford, K. H. 1992. The empty forest. BioScience 42: and outcrossing plant species. Ecology and Evolution
412–422. (2, 4) 5(23): 5642–5651. (10)
Redford, K. H. and W. M. Adams. 2009. Payment for eco- Reiter, N., J. Whitfield, G. Pollard, W. Bedggood, et al.
system services and the challenge of saving nature. 2016. Orchid re-introductions: an evaluation of
Conservation Biology 23: 785–787. (3) success and ecological considerations using key
comparative studies from Australia. Plant Ecology
Redford, K. H., G. Amato, J. Baillie, P. Beldomenico, et al.
1–15. (7)
2011. What does it mean to successfully conserve a
(vertebrate) species? BioScience 61: 39–48. (6) Reyers, B., R. Biggs, G. S. Cumming, T. Elmqvist, et al.
2013. Getting the measure of ecosystem services: a
Redford, K. H., B. J. Huntley, D. Roe, T. Hammond, et al.
social-ecological approach. Frontiers in Ecology and
2015. Mainstreaming Biodiversity: Conservation for
the Environment 11: 268–273. (3)
the Twenty-First Century. Frontiers in Ecology and
Evolution 3: 137. (12) Reyers, B., D. J. Roux, R. M. Cowling, A. E. Ginsburg,
et al. 2010. Conservation planning as a transdisci-
Redpath, S. M., J. Young, A. Evely, W. M. Adams, et al.
plinary process. Conservation Biology 24: 957–965. (1)
2013. Understanding and managing conservation
conflicts. Trends in Ecology and Evolution 28: 100–109. Ribeiro, J., G. R. Colli, J. P. Caldwell, and A. M. Soares.
(9, 12) 2016. An integrated trait-based framework to pre-
dict extinction risk and guide conservation planning
Reed, D. H., J. J. O’Grady, J. D. Ballou, and R. Frankham.
2003. The frequency and severity of catastrophic
452 Bibliography

in biodiversity hotspots. Biological Conservation 195: Robertson, M. M. 2006. Emerging ecosystem service mar-
214–223. (9) kets: Trends in a decade of entrepreneurial wetland
Ricciardi, A. 2003. Predicting the impacts of an intro- banking. Frontiers in Ecology and the Environment 4:
duced species from its invasion history: An empiri- 297–302. (10)
cal approach applied to zebra mussel invasions. Robinson, J. G. 2011. Ethical pluralism, pragmatism, and
Freshwater Biology 48: 972–981. (4) sustainability in conservation practice. Biological
Richardson, C. J. and N. J. Hussain. 2006. Restoring the Conservation 144: 958–965. (3)
Garden of Eden: an ecological assessment of the Robinson, J. G. 2012. Common and conflicting interests
marshes of Iraq. BioScience 56: 477–489. (10) in the engagements between conservation organi-
Richardson, S. J., R. Clayton, B. D. Rance, H. Broadbent, zations and corporations. Conservation Biology 26:
et al. 2015. Small wetlands are critical for safe- 967–977. (11, 12)
guarding rare and threatened plant species. Applied Robinson, R. A., H. P. Q. Crick, J. A. Learmonth, I. M. D.
Vegetation Science 18(2): 230–241. (8) Maclean, et al. 2008. Travelling through a warm-
Ricketts, T. H., E. Dinerstein, D. M. Olson, C. J. Loucks, ing world: Climate change and migratory species.
et al. 1999. Terrestrial Ecoregions of North America: A Endangered Species Research 7: 87–99. (4)
Conservation Assessment. Island Press, Washington, Robles, M. D., C. H. Flather, S. M Stein, M. D. Nelson,
DC. (2) and A. Cutko. 2008. The geography of private for-
Ricklefs, R. E., and F. He. 2016. Region effects influence ests that support at-risk species in the conterminous
local tree species diversity. Proceedings of the National United States. Frontiers in Ecology and the Environ-
Academy of Sciences USA, 113: 674–679. (2) ment 6: 301–307. (9)
Riehl, B., H. Zerriffi, and R. Naidoo. 2015. Effects of Rochefort, L. and E. Lode. 2006. Restoration of de-
community-based natural resource management graded boreal peatlands. In R. K. Wieder and D. H.
on household welfare in Namibia. PLoS ONE 10(5): Vitt (eds.), Boreal Peatland Ecosystems, p. 381–423.
e0125531. (9) Springer Berlin Heidelberg. (10)
Rife, A. N., B. Erisman, A. Sanchez, and O. Aburto- Roe, D., E. Y. Mohammed, I. Porras, and A. Giuliani.
Oropeza. 2013. When good intentions are not 2013. Linking biodiversity conservation and poverty
enough … Insights on networks of “paper park” reduction: De-polarizing the conservation–poverty
marine protected areas. Conservation Letters 6: debate. Conservation Letters 6: 162–171. (9, 11)
200–212. (8) Rolston, H., III. 1994. 2012. Environmental Ethics. Temple
Rinella, M. F., B. D. Maxwell, P. K. Fay, T. Weaver, and R. University Press. (3)
L. Sheley. 2009. Control effort exacerbates invasive- Roman, J., Dunphy-Daly, D. W. Johnston, and A. J. Read.
species problem. Ecological Applications 19: 155–162. 2015. Lifting baselines to address the consequences
(4) of conservation success. Trends in Ecology and Evolu-
Ripple, W. J. and R. L. Beschta. 2012. Trophic cascades in tion 30.6: 299–302. (1)
Yellowstone: The first 15 years after wolf reintro- Roman, J. and S. R. Palumbi. 2003. Whales before whaling
duction. Biological Conservation 145: 205–213. (2) in the North Atlantic. Science 301(5632): 508–510. (4)
Ripple, W. J., J. A. Estes, R. L. Beschta, C. C. Wilmers, Romero, G. Q., T. Gonçalves-Souza, C. Vieira, and J.
et al. 2014. Status and ecological effects of the Koricheva. 2015. Ecosystem engineering effects on
world’s largest carnivores. Science 343: 1241484. doi: species diversity across ecosystems: a meta-analysis.
10.1126/science1241484. (2, 5, 11) Biological Reviews 90(3): 877–890. (2)
Riskas, K. A., M. M. Fuentes, and M. Hamann. 2016) Rompré, G., W. D. Robinson, A. Desrochers, and G.
Justifying the need for collaborative management Angehr. 2009. Predicting declines in avian species
of fisheries bycatch: a lesson from marine turtles in richness under nonrandom patterns of habitat loss
Australia. Biological Conservation 196: 40–47. in a Neotropical landscape. Ecological Applications
Rissman, A. R. and V. Butsic. 2011. Land trust defense 19: 1614–1627. (5)
and enforcement of conserved areas. Conservation Rood, S. B., S. Kaluthota, K. M. Gill, E. J. Hillman, et al.
Letters 4: 31–37. (11) 2015. A twofold strategy for riparian restoration:
Rissman, A. R., J. Owley, M. R. Shaw, and B. B. Thomp- combining a functional flow regime and direct seed-
son. 2015. Adapting conservation easements to ing to re-establish cottonwoods. River Research and
climate change. Conservation Letters 8(1): 68–76. (11) Applications. doi: 10.1002/rra.2919. (7)
Riverá Ortíz, F. A., R. Aguilar, M. D. C. Arizmendi, M. Rosa, I. B. D., C. Souza Jr., and R. M. Ewers. 2012.
Quesada, and K. Oyama. 2014. Habitat fragmenta- Changes in size of deforested patches in the Brazil-
tion and genetic variability of tetrapod populations. ian Amazon. Conservation Biology 26: 932–937. (4)
Animal Conservation 18(3): 249–258. (5) Roscher, C., J. Schumacher, B. Schmid, E-D. Schulze.
Robertson, G. C. Moreno, J. A. Arata, S. G. Candy, et al. 2015. Contrasting effects of intraspecific trait varia-
2014. Black-browed albatross numbers in Chile in- tion on trait-based niches and performance of le-
crease in response to reduced mortality in fisheries. gumes in plant miztures. PloS ONE 10(3) e0119786.
Biological Conservation 169: 319–333. (4) doi: 10.1371/journal. pone.0119786. (2)
Bibliography 453

Rosenzweig, M. L. 2003. Win-win ecology: how the Earths Sandom, C., S. Faurby, B. Sandel, and J. C. Svenning.
species can survive in the midst of human enter- 2014. Global late Quaternary megafauna extinctions
prise. Oxford University Press, New York. (9, 10) linked to humans, not climate change. Proceedings
Ruane, J. 2000. A framework for prioritizing domestic of the Royal Society of London B: Biological Sciences
animal breeds for conservation purposes at the 281(1787): 20133254. (5)
national level: A Norwegian case study. Conservation Sanford, E., B. Gaylord, A. Hettinger, E. A. Lenz, et al.
Biology 14: 1385–1393. (7) 2014. Ocean acidification increases the vulnerability
Ruscoe, W. A., P. J. Sweetapple, M. Perry, and R. P. Dun- of native oysters to predation by invasive snails.
can. 2013. Effects of spatially extensive control of Proceedings of the Royal Society of London B: Biological
invasive rats on abundance of native invertabrates Sciences 281(1778): 20132681. (4)
in mainland New Zealand forests. Conservation Biol- Sanjayan, M., L. H. Samberg, T. Boucher, and J. Newby.
ogy 27: 74–82. (4) 2012. Intact faunal assemblages in the modern era.
Russell, J. C., J. G. Innes, P. H. Brown, A. E. Byrom. 2015a. Conservation Biology 26: 724–730. (5)
Predator-free New Zealand: conservation country. Sauer, J. R., J. E. Hines, J. E. Fallon, K. L. Pardieck, et al.
BioScience, biv012. (4) 2014. The North American Breeding Bird Survey,
Russell, J. C., H. P. Jones, D. P. Armstrong, F. Courchamp, Results and Analysis 1966–2013. Version 01.30.2015.
et al. 2015b. Importance of lethal control of invasive USGS Patuxent Wildlife Research Center, Laurel,
predators for island conservation. Conservation Biol- MD. (6)
ogy. (4) Sauer, J. R., W. A. Link, J. E. Fallon, K. L. Pardieck, and
Rust, N. A. 2015. Media Framing of Financial Mecha- D. J. Ziolkowski, Jr. 2013. The North American
nisms for Resolving Human–Predator Conflict breeding bird survey 1966-2011: summary analysis
in Namibia. Human Dimensions of Wildlife 20(5): and species accounts. North American Fauna 79(79):
440–453. (3) 1–32. (6)
Ryan, M. E. J. R. Johnson, B. M. Fitzpatrick, L. J. Lowen- Sayre, N. F., R. R. McAllister, B. T. Bestelmeyer, M.
stine, et al. 2013. Lethal effects of water quality on Moritz, et al. 2013. Earth Stewardship of rangelands:
threatened California salamanders but not on co- coping with ecological, economic, and political
occurring hybrid salamanders. Conservation Biology marginality. Frontiers in Ecology and the Environment
27: 95–102. (2) 11: 348–354. (4, 12)
Scariot, A. 2013. Land sparing or land sharing: the miss-
ing link. Frontiers in Ecology and the Environment 11:
Saarinen, K., A. Valtonen, J. Jantunen, and S. Saarino.
177–178. (9)
2005. Butterflies and diurnal moths along road verg-
es: does road type affect diversity and abundance? Schaider, L. A., R. A. Rudel, J. M. Ackerman, S. C.
Biological Conservation 123: 403–412. (9) Dunagan, and J. G. Brody. 2014. Pharmaceuticals,
perfluorosurfactants, and other organic wastewater
Saccheri, I., M. Kuussaari, M. Kankare, P. Vikman, et al.
compounds in public drinking water wells in a
1998. Inbreeding and extinction in a butterfly meta-
shallow sand and gravel aquifer. Science of the Total
population. Nature 392: (6675): 491–494. (5)
Environment 468: 384–393. (4)
Sachs, J. D. 2008. Common Wealth: Economics for a Crowded
Scheckenbach, F., K. Hausmann, C. Wylezich, M. Weitere,
Planet. Penguin Press, New York. (3)
and H. Arndt. 2010. Large-scale patterns in biodi-
Sæther, B. E. and S. Engen. 2015. The concept of fitness versity of microbial eukaryotes from the abyssal sea
in fluctuating environments. Trends in Ecology and floor. Proceedings of the National Academy of Sciences
Evolution 30(5): 273–281. (6) USA 107: 115–120. (2)
Sagoff, M. 2008. On the compatibility of a conservation Scheffer, M., S. Barrett, S. R. Carpenter, C. Folke, C., et
ethic with biological science. Conservation Biology 21: al. 2015. Creating a safe operating space for iconic
337–345. (3) ecosystems. Science 347(6228): 1317–1319. (8)
Sairam, R., S. Chennareddy, and M. Parani. 2005. OBPC Schlacher, T. A., and L. Thompson. 2012. Beach recreation
Symposium: Maize 2004 and Beyond—plant regen- impacts benthic invertebrates on ocean-exposed
eration, gene discovery, and genetic engineering of sandy shores. Biological Conservation 147: 123–132. (3)
plants for crop improvement. In Vitro Cellular and
Schleuning, M. and D. Matthies. 2009. Habitat change
Developmental Biology—Plant 41: 411. (3)
and plant demography: Assessing the extinction
Sajeva, M., C. Augugliaro, M. J. Smith, and E. Oddo. risk of a formerly common grassland perennial.
2013. Regulating internet trade in CITES species. Conservation Biology 23: 174–183. (5)
Conservation Biology 27: 429–430. (6)
Schmidt, C., Kucera, M., and Uthicke, S. 2014. Combined
Sale, P. F. 2015. Coral reef conservation and political will. effects of warming and ocean acidification on coral
Environmental Conservation 42(02): 97–101. (12) reef Foraminifera Marginopora vertebralis and Hetero-
Sanderson, E., M. Jaiteh, M. A. Levy, K. H. Redford, et al. stegina depressa. Coral Reefs 33(3): 805–818. (4)
2002. The human footprint and the last of the wild. Schmitz, P., S. Caspers, P. Warren, P., and K. Witte. 2015.
BioScience 52: 891–904. (4) First steps into the wild–exploration behavior of
454 Bibliography

European bison after the first reintroduction in Şekercioğlu, Ç. H., R. B. Primack, and J. Wormworth.
Western Europe. PLoS ONE 10(11). (7) 2012. The effects of climate change on tropical birds.
Schofield, G., R. Scott, A. Dimadi, S. Fossette, et al. 2013. Biological Conservation 148(1): 1–18. (4, 5)
Evidence-based marine protected area planning Selier, S. A. J., B. R. Page, A. T. Vanak, and R. Slotow.
for a highly mobile endangered marine vertebrate. 2014. Sustainability of elephant hunting across in-
Biological Convservation. 161: 101–109. (8) ternational borders in southern Africa: A case study
Schonewald-Cox, C. M. 1983. Conclusions: guidelines of the greater Mapungubwe Transfrontier Conser-
to management: a beginning attempt. In C. M. vation Area. The Journal of Wildlife Management 78(1):
Schonewald-Cox, S. M. Chambers, B. MacBryde and 122–132. (3)
L. Thomas (eds.), Genetics and Conservation: A Refer- Selomane, O, B. Reyers, R. Biggs, H. Tallis, and S.
ence for Managing Wild Animal and Plant Populations, Polasky. 2015. Towards integrated social-ecological
pp. 414–445. Benjamin/Cummings, Menlo Park, sustainability indicators: Exploring the contribution
CA. (6) and gaps in existing global data. Ecological Econom-
Schuur, E. A. G., A. D. McGuire, C. Schädel, G. Grosse, et ics 118: 140–146. (11)
al. 2015. Climate change and the permafrost carbon Sethi, P. and H. F. Howe. 2009. Recruitment of hornbill-
feedback. Nature 520(7546): 71–179. (4) dispersed trees in hunted and logged forests of the
Schwartz, M. W. 2008. The performance of the Endan- eastern Indian Himalaya. Conservation Biology 23:
gered Species Act. Annual Review of Ecology, Evolu- 710–718. (3)
tion, and Systematics 39: 279–299. (6) Shackelford, N., R. J. Hobbs, N. E. Heller, L. M. Hallett, et
Schwenk, W. S. and T. M. Donovan. 2011. A multispecies al. 2013. Finding a middle-ground: The native/non-
framework for landscape conservation planning. native debate. Biological Conservation 158: 55–62. (4)
Conservation Biology 25: 1010–1021. (8) Shafer, C. L. 1997. Terrestrial nature reserve design at
Schwitzer, C., R. A. Mittermeier, S. E. Johnson, G. Donati, the urban/rural interface. In M. W. Schwartz (ed.),
et al. 2014. Averting lemur extinctions amid Mada- Conservation in Highly Fragmented Landscapes, pp.
gascar’s political crisis. Science 343: 842–843. (5) 345–378. Chapman and Hall, New York. (8)
Scott, J. M., B. Csuti, and F. Davis. 1991. Gap analysis: An Shafer, C. L. 1999. History of selection and system
application of Geographic Information Systems for planning for U. S. natural area national parks and
wildlife species. In D. J. Decker, M. E. Krasny, G. R. monuments: beauty and biology. Biodiversity and
Goff, C. R. Smith, and D. W. Gross (eds.), Challenges Conservation 8: 189–204. (8)
in the Conservation of Biological Resources: A Practitio- Shafer, C. L. 2001. Conservation biology trailblazers:
ner’s Guide, pp. 167–179. Westview Press, Boulder, George Wright, Ben Thompson, and Joseph Dixon.
CO. (8) Conservation Biology 15: 332–344. (1)
Scott, J. M., F. W. Davis, R. G. McGhie, R. G. Wright, et Shaffer, M. L. 1981. Minimum population sizes for spe-
al. 2001. Nature reserves: do they capture the full cies conservation. BioScience 31: 131–134. (6)
range of America’s biological diversity? Ambio 11: Shafroth, P. B., V. B. Beauchamp, M. K. Briggs, K. Lair, et
999–1007. (8) al. 2008. Planning riparian restoration in the context
Scott, J. M., D. D. Goble, J. A. Wiens, D. S. Wilcove, et of Tamarix control in western North America. Resto-
al. 2005. Recovery of imperiled species under the ration Ecology, 16(1): 97–112. (10)
Endangered Species Act: The need for a new ap- Shanley, P. and L. Luz. 2003. The impacts of forest deg-
proach. Frontiers in Ecology and the Environment 3: radation on medicinal plant use and implications
383–389. (6) for health care in eastern Amazonia. BioScience 53:
Scott, J. M., F. L. Ramse, M. Lammertink, K. V. Rosen- 573–584. (3)
berg, et al. 2008. When is an “extinct” species really Sharma, N., M. D. Madhusudan, and A. Sinha. 2014. Lo-
extinct? Gauging the search efforts for Hawaiian cal and landscape correlates of primate distribution
forest birds and the ivory-billed woodpecker. Avian and persistence in the remnant lowland forests of
Conservation and Ecology 3(2): 3. (5) the Upper Brahmaputra Valley, northeastern India.
Seastedt, T. R. 2015. Biological control of invasive plant Conservation Biology 28: 95–106. (5)
species: a reassessment for the Anthropocene. New Sher, A. Introduction to the Paradox Plant. 2013. In A.
Phytologist 205(2): 490–502. (10) Sher and M. F. Quigley (eds.), Tamarix: A Case Study
Sebastián-González, E., J. A. Sánchez-Zapata, F. Botella, J. of Ecological Change in the American West, pp. 1–20.
Figuerola, et al. 2011. Linking cost efficiency evalu- Oxford University Press, New York. (10)
ation with population viability analysis to prioritize Sher, A. A. and L. A. Hyatt. 1999. The disturbed resource-
wetland bird conservation actions. Biological Conser- flux invasion matrix: a new framework for pat-
vation 144: 2354–2361. (7) terns of plant invasion. Biological Invasions 1(2–3):
Seddon, P. J., C. J. Griffiths, P. S. Soorae, and D. P. 107–114. (4)
Armstrong. 2014. Reversing defaunation: Restor- Shutt, K., M. Heistermann, A. Kasim, A. Todd, et al. 2014.
ing species in a changing world. Science 345(6195): Effects of habituation, research and ecotourism on
406–412. (7) faecal glucocorticoid metabolites in wild western
Bibliography 455

lowland gorillas: implications for ecotourism. Bio- Sogge, M. K, E. H. Paxton, C. Van Riper. 2013. Tamarix in
logical Conservation 172: 72–79. (3) riparian woodlands: a bird’s eye view. In A. Sher
Shwartz, A., A. Turbé, L. Simon, R. Juliard. 2014. Enhanc- and M. F. Quigley (eds.), Tamarix: A Case Study of
ing urban biodiversity and its influence on city Ecological Change in the American West, pp. 189–206.
dwellers. Biological Conservation 171: 82–90. (10) Oxford University Press, New York. (10)
Siikamäki, J. 2011. Contributions of the US state park sys- Solow, A., W. Smith, M. Burgman, T. Rout, T. et al.
tem to nature recreation. Proceedings of the National 2012. Uncertain sightings and the extinction of the
Academy of Sciences USA 108: 14031–14036. (3) ivory-billed woodpecker. Conservation Biology 26(1):
Simaika, J. P., M. J. Samways, J. Kipping, F. Suhling, et al. 180–184. (5)
2013. Continental-scale conservation prioritization Sorice, M. G., C-O. Oh, T. Gartner, M. Snieckus, et al.
of African dragonflies. Biological Conservation 157: 2013. Increasing participation in incentive programs
245–254. (2, 8) for biodiversity conservation. Ecological Applications
Simberloff, D. S., J. A. Farr, J. Cox, and D. W. Mehlman. 23: 1146–1155. (11)
1992. Movement corridors: conservation bargains or Soulé, M. E. 1985. What is conservation biology? BioSci-
poor investments? Conservation Biology 6: 493–505. (8) ence 35: 727–734. (1)
Simberloff, D., C. Murcia, and J. Aronson, J. 2015. Novel Soulé, M. E. and D. Simberloff. 1986. What do genet-
ecosystems are a Trojan horse for conservation. ics and ecology tell us about the design of nature
Ensia. ensia.com/voices/novel-ecosystems-are-a-trojan- reserves? Biological Conservation 35: 19–40. (8)
horse-for-conservation. (10) Spalding, M. D., L. Fish, and L. J. Wood. 2008. Towards
Simmons, R. E. 1996. Population declines, variable breed- representative protection of the world’s coasts and
ing areas and management options for flamingos in oceans—Progress, gaps, and opportunities. Conser-
Southern Africa. Conservation Biology 10: 504–515. (6) vation Letters 1: 217–226. (8)
Siraj, A. S., M. Santos-Vega, M. Bouma, D. Yadeta, et al. Spielman, D., B. W. Brook, and R. Frankham. 2004. Most
2014. Altitudinal changes in malaria incidence in species are not driven to extinction before genetic
highlands of Ethiopia and Colombia. Science 343: factors impact them. Proceedings of the National Acad-
1154–1158. (4) emy of Sciences USA 101: 15261–15264. (5)
Skelly, D. K., S. R. Bolden, and L. K. Freidenburg. 2014. Sponberg, A. F. 2009. Great Lakes: Sailing to the forefront
Experimental canopy removal enhances diversity of of national water policy? BioScience 59: 372. (10)
vernal pool amphibians. Ecological Applications 24: Srinivasan, U. T., S. P. Carey, E. Hallstein, P. A. T. Higgins,
340–345. (8) et al. 2008. The debt of nations and the distribu-
Smith, A. 1909. An Inquiry into the Nature and Causes of tion of ecological impacts from human activities.
the Wealth of Nations. In J. L. Bullock (ed.), The Har- Proceedings of the National Academy of Sciences USA
vard Classics, P. F. Collier and Sons, New York. (3) 105: 1768–1773. (6)
Smith, S., A. A. Sher, and T. A. Grant. 2007. Genetic Diver- Stankey, G. H. and B. Shindler. 2006. Formation of social
sity in Restoration Materials and the Impacts of Seed acceptability judgments and their implications for
Collection in Colorado’s Restoration Plant Produc- management of rare and little-known species. Con-
tion Industry. Restoration Ecology 15: 369–374. (10) servation Biology 20(1): 28–37. (6)
Smyser, T. J., S. A. Johnson, K. Page, C. M. Hudson, et al. Staricco, J. I. and S. Ponte. 2015. Quality regimes in agro-
2013. Use of experimental translocations of allegh- food industries: A regulation theory reading of Fair
eny woodrat to decipher casual agents of decline. Trade wine in Argentina. Journal of Rural Studies
Conservation Biology, 27: 752–762. (7) 66–76. (12)
Soanes, K., M. C. Lobo, P. A. Vesk, M. A. McCarthy, et al. Steadman, D. W. 1995. Prehistoric extinctions of Pacific
2013. Movement re-established but not restored: in- island birds: biodiversity meets zooarchaeology.
ferring the effectiveness of road-crossing mitigation Science 267(5201: 1123–1131. (5)
for a gliding mammal by monitoring use. Biological Stein, B. A., L. S. Kutner, and J. S. Adams (eds.). 2000. Pre-
Conservation 159: 434–441. (8) cious Heritage: The Status of Biodiversity in the United
Soares-Filho, B., P. Moutinho, D. Nepstad, A. Anderson, States. Oxford University Press, New York. (4)
et al. 2010. Role of Brazilian Amazon protected areas Stein, B. A., C. Scott, and N. Benton. 2008. Federal lands
in climate change mitigation. Proceedings of the Na- and endangered species: the role of the military
tional Academy of Sciences USA 107: 10821–10826. (9) and other federal lands in sustaining biodiversity.
Society for Conservation Biology. 2016. What is SCB? BioScience 58: 339–347. (9)
Organizational Values. conbio.org/about-scb/who-we- Stocks, A. 2005. Too much for too few: problems of indig-
are (1) enous land rights in Latin America. Annual Review
Society for Ecological Restoration. www.ser.org (10) of Anthropology 34: 85–104. (12)
Sodhi, S. N., R. Butler, W. F. Laurance, and L. Gibson. Stokes, E. J., S. Strindberg, P. C. Bakabana, P. W. Elkan, et
2011. Conservation successes at micro-, meso- and al. 2010. Monitoring great ape and elephant abun-
macroscales. Trends in Ecology and Evolution 26: dance at large spatial scales: Measuring effective-
585–594. (1) ness of a conservation landscape. PLoS ONE 5(4):
e10294. (9)
456 Bibliography

Stokstad, E. 2009. Obama moves to revitalize Chesapeake Tallis, H. and S. Polasky. 2009. Mapping and valuing
Bay restoration. Science 324: 1138–1139. (10) ecosystem services as an approach for conservation
Stouffer, P. C., E. I. Johnson, R. O. Bierregaard Jr, and T. and natural-resource management. Annals of the
E. Lovejoy. 2011. Understory bird communities in New York Academy of Sciences 1162: 265–283. (6)
Amazonian rainforest fragments: Species turnover Tamarisk Coalition. 2016. Tamarisk beelte (Diorhabda
through 25 Years post-isolation in recovering land- spp.) in the Colorado River basin: synthesis of an
scapes. PLoS ONE 6(6): e20543. (4) expert panel forum. Tamarisk Coalition, Grand
Strain, D. 2011. 8.7 million: A new estimate for all the Junction, CO. (10)
complex species on earth. Science 333: 1083. (2) Taylor, M. F. J., K. F. Suckling, and J. J. Rachlinski. 2005.
Strayer, D. L. 2009. Twenty years of zebra mussels: The effectiveness of the Endangered Species Act: A
lessons from the mollusk that made headlines. Fron- quantitative analysis. BioScience 55: 360–366. (6)
tiers in Ecology and the Environment 7: 135–141. (4) Temperton, V. M. and R. J. Hobbs. 2004. The search
Sullivan, B. L., J. L. Aycrigg, J. H. Barry, R. E. Bonney, et for ecological assembly rules and its relevance to
al. 2014. The eBird enterprise: An integrated ap- restoration ecology. In V. M. Temperton, R. J. Hobbs,
proach to development and application of citizen T. J. Nuttle, and S Halle (eds.), Assembly Rules And
science. Biological Science 169: 31–40. (12) Restoration Ecology—Bridging The Gap Between Theory
And Practice, pp. 34–54. Island Press, Washington,
Sun, Z., Q. Huang, C. Opp, T. Hennig, T., and U. Marold.
DC. (10)
2012. Impacts and implications of major changes
caused by the Three Gorges Dam in the middle Temple, S. A. 1991. Conservation biology: New goals and
reaches of the Yangtze River, China. Water Resources new partners for managers of biological resources.
Management 26(12): 3367–3378. (4) In D. J. Decker, M. Krasny, G. R. Goff, C. R. Smith,
and D. W. Gross (eds.), Challenges in the Conserva-
Sutherland, W. J., W. M. Adams, R. B. Aronson, R.
tion of Biological Resources: A Practitioner’s Guide, pp.
Aveling, et al. 2009. One hundred questions of
45–54. Westview Press, Boulder, CO. (1)
importance to the conservation of global biological
diversity. Conservation Biology 23: 557–567. (12) Templeton, A. R. 1986. Coadaptation and outbreeding
depression. In M. E. Soulé (ed.), Conservation Biol-
Sutherland, W. J., S. Bardsley, L. Bennun, M. Clout, et al.
ogy: The Science of Scarcity and Diversity, pp. 105–116.
2011. Horizon scan of global conservation issues for
Sinauer Associates, Sunderland, MA. (5)
2011. Trends in Ecology & Evolution 26: 10–16. (12)
Tende, T., B. Hansson, U. Ottosson, M. Åkesson, and S.
Svenning, J. C., P. B. Pedersen, C. J. Donlan, R. Ejrnæs,
Bensch. 2014. Individual identification and genetic
et al. 2015. Science for a wilder Anthropocene:
variation of lions (Panthera leo) from two protected
Synthesis and future directions for trophic rewild-
areas in Nigeria. PLoS ONE 9(1): e84288. (5)
ing research. Proceedings of the National Academy of
Sciences 113(4): 898–906. (10) Thatcher, C. A., F. T. van Manen, and J. D. Clark. 2009. A
habitat assessment for Florida panther population
Swanson, F. J., C. Goodrich, and K. D. Moore. 2008.
expansion into central Florida. Journal of Mammalogy
Bridging boundaries: scientists, creative writers,
90: 918–925. (9)
and the long view of the forest. Frontiers in Ecology
and the Environment 6: 499–504. (3, 12) Theobald, D. M. 2004. Placing exurban land-use change
in a human modification framework. Frontiers in
Sweka, J. A. and T. C. Wainwright. 2014. Use of popu-
Ecology and the Environment 2: 139–144. (9)
lation viability analysis models for Atlantic and
Pacific salmon recovery planning. Reviews in Fish Thiollay, J. M. 1989. Area requirements for the conserva-
Biology and Fisheries 24(3): 901–917. (6) tion of rainforest raptors and game birds in French
Guiana. Conservation Biology 3: 128–137. (6)
Szabo, J. K., S. H. M. Butchart, H. P. Possingham, and S. T.
Garnett. 2012. Adapting global biodiversity indica- Thogmartin, W. E., C. A. Sander-Reed, J. A. Szymanski,
tors to the national scale: a Red List index for Aus- P. C. McKann, et al. 2013. White-nose syndrome is
tralian birds. Biological Conservation 148: 61–68. (6) likely to extirpate the endangered Indiana bat over
large parts of its range. Biological Conservation 160:
Szlávik, J. and M. Füle. 2009. Economic consequences of
162–171. (4)
climate change. American Institute of Physics Con-
ference Proceedings, Sustainability 2009: The Next Thomsen P. F. and W. Willerslev. 2015. Environmen-
Horizon 1157: 73–82. (12) tal DNA—an emerging tool in conservation for
monitoring past and present biodiversity. Biological
Conservation 183: 4–18. (6)
Tallis, H., R. Goldman, M. Uhl, and B. Brosi. 2009. Inte-
Thondhlana, G. S. Shackleton, and J. Blignaut. 2015.
grating conservation and development in the field:
Local institutions, actors, and natural resource
implementing ecosystem service projects. Frontiers
governance in Kgalagadi Transfrontier Park and
in Ecology and the Environment 7: 12–20. (9)
surrounds, South Africa. Land Use Policy 47: 121–129
Tallis, H., J. Lubchenco, V. M. Adams, C. Adams-Hosk- doi: 10.1016/j.landusepol.2015.03.013. (11)
ing, et al. 2014. Working together: A call for inclu-
Thoreau, H. D. 1854. Walden; or, Life in the Woods. Ticknor
sive conservation. Nature 515(7525): 27–28. (12)
and Fields, Boston. (3)
Bibliography 457

Thorp, J. H., J. E. Flotemersch, M. D. Delong, A. F. Turner, I. M. 2001. The Ecology of Trees in the Tropical Rain
Casper, et al. 2010. Linking ecosystem services, Forest. Cambridge University Press. (4)
rehabilitation, and river hydrogeomorphology. Turvey, S. 2008. Witness to Extinction: How We Failed to
BioScience 60: 67–74. (3) Save the Yangtze River Dolphin. Oxford University
Tilman, D., R. M. May, C. L. Lehman, and M. A. Nowak, Press. (5)
M. A. 1994. Habitat destruction and the extinction Twilley, R. R. and J. W. Day. 2012. Mangrove wetlands. In
debt. Nature 371: 65–66. (5) J. W. Day et al. (ed.), Estuarine Ecology, 2nd Ed., pp.
Timmer, V. and C. Juma. 2005. Biodiversity conservation 165–202. Wiley-Blackwell, Hoboken, NJ. (4)
and poverty reduction come together in the tropics:
Lessons learned from the Equator Initiative. Envi- U. S. Census Bureau. www.census.gov (1)
ronment 47: 25–44. (9)
U. S. Energy Information Administration. 2015. Interna-
Tingley, R., R. A. Hitchmough, and D. G. Chapple. 2013. tional Energy Statistics 2008–2011. www.eia.gov/tools/
Life-history traits and extrinsic threats determine faqs/faq.cfm?id=85&t=1 (1)
extinction risk in New Zealand lizards. Biological
Conservation 165: 62–67. (5) U. S. Fish and Wildlife Service (USFWS). 2003. Guidance
for the Establishment, Use, and Operation of Con-
Tittensor, D. P., C. Mora, W. Jetz, H. K. Lotze, et al. 2010. servation Banks. www.fws.gov/endangered/esa-library/
Global patterns and predictors of marine biodiver- pdf/Conservation_Banking_Guidance.pdf (11)
sity across taxa. Nature 466: 1098–1101. (2)
U. S. Fish and Wildlife Service (USFWS). 2013. Federal
Tlusty, M. F., A. L. Rhyne, L. Kaufman, M. Hutchins, et and State Endangered and Threatened Species Ex-
al. 2013. Opportunities for public aquariums to in- penditures Fiscal Year 2013. www.fws.gov/endangered/
crease the sustainability of the aquatic animal trade. esa-library/pdf/2012.EXP.FINAL.pdf, p. 5. (6)
Zoo Biology 32: 1–12. (7)
U. S. National Park Service. 2009. National Park of
Tognelli, M. F., M. Fernández, and P. A. Marquet. 2009. American Samoa. www.nps.gov/npsa/index.htm (9)
Assessing the performance of the existing and
proposed network of marine protected areas to U. S. National Park Service. www.nps.gov (9)
conserve marine biodiversity in Chile. Biological UNESCO World Heritage Centre. 2010. whc.unesco.org (11)
Conservation 142: 3147–3153. (8) United Nations Development Programme. 2006. www.
Toledo, V. M. 2001. Indigenous peoples and biodiversity. undp.org (4)
In S. A. Levin (ed.), Encyclopedia of Biodiversity 3: United Nations Environment Programme (UNEP). 2014.
451–464. Academic Press, San Diego, CA. (9) Protected Planet Report. www.unep-wcmc.org/resourc-
Towne, E. G., D. C. Hartnett, and R. C. Cochran. 2005. Veg- es-and-data/protected-planet-report-2014 (8)
etation trends in tallgrass prairie from bison and cattle United Nations Environment Programme World Con-
grazing. Ecological Applications 15: 1550–1559. (8) servation Monitoring Centre (UNEP-WCMC). 2014.
Traill, L. W., B. W. Brook, R. R. Frankham, and C. J. A. Report on achievements for the year 2014. Cam-
Bradshaw. 2010. Pragmatic population viability bridge, UK. (8)
targets in a rapidly changing world. Biological Con- United Nations Environment Programme Mediterranean
servation 143: 28–34. (6) Action Plan (UNEPMAP). 2016. Countries Agree
Tran, P. M. and L. Waller. 2013. Effects of landscape frag- Ambitious Conservation Measures for Mediterra-
mentation and climate on Lyme disease incidence nean. www.unep.org/Documents.Multilingual/Default.
in the northeastern United States. Ecohealth 10(4): asp?DocumentID=27058&ArticleID=35947&l=en (9)
394–404. (4) United Nations. 1993a. Agenda 21: Rio Declaration and
Tranquilli, S., M. Abedi-Lartey, F. Amsini, L. Arranz, et al. Forest Principles. Post-Rio Edition. United Nations
2012. Lack of conservation effort rapidly increases Publications, New York. (11)
African great ape extinction risk. Conservation Letters United Nations. 1993b. The Global Partnership for Environ-
5: 48–55. (8) ment and Development. United Nations Publications,
Tree of Life Web Project. tolweb.org (12) New York. (11)
Triantis, K., D. Nogues-Bravo, J. Hortal, A. V. Borges, Uriarte, M., M. Pinedo-Vasquez, R. S. DeFries, K.
et al. 2008. Measurements of area and the (island) Fernandes, et al. 2012. Depopulation of rural
species-area relationship: new directions for an old landscapes exacerbates fire activity in the western
pattern. Oikos 117: 1555–1559. (5) Amazon. Proceedings of the National Academy of Sci-
Troupin, D. and Y. Carmel. 2014. Can agro-ecosystems ences 109(52): 21546–21550. (4)
efficiently complement protected area networks?
Biological Conservation 169: 158–166. (9) Valiela, I. and P. Martinetto. 2007. Changes in bird abun-
Tscharntke, T., Y. Clough, T. C. Wanger, L. Jackson, et al. dance in eastern North America: Urban sprawl and
2012. Global food security, biodiversity conserva- global footprint? BioScience 57: 360–370. (4)
tion and the future of agricultural intensification. van Swaay, C., D. Maes, S. Collins, M. L. Munguira, et al.
Biological Conservation 151: 53–59. (9) 2011. Applying IUCN criteria to inverte-brates: How
red is the Red List of European butterflies? Biological
Conservation 144: 470–478. (6)
458 Bibliography

Van Teeffelen, A. J. A., C. C. Vos, and P. Opdam. 2012. Vredenburg, V. T. 2004. Reversing introduced species ef-
Species in a dynamic world: consequences of habi- fects: Experimental removal of introduced fish leads
tat network dynamics on conservation planning. to rapid recovery of a declining frog. Proceedings of the
Biological Conservation 153: 239–253. (8) National Academy of Sciences USA 101: 7646–7650. (4)
Van Vugt, M. 2009. Averting the tragedy of the commons:
using social psychological science to protect the en- Wackernagel, M. and W. Rees. 1996. Our Ecological
vironment. Current Directions in Psychological Science Footprint: Reducing Human Impact on the Earth (New
18: 169–173. (11) Catalyst. Bioregional Series). New Society Publishers,
Van Wilgen, B. W., G. G. Forsyth, D. C. Le Maitre, A. Gabriola Island, BC. (4)
Wannenburgh, et al. 2012. An assessment of the Wade, L. 2013. Gold’s dark side. Science 341: 1448–1449. (4)
effictiveness of a large, national-scale invasive alien Wagner, K. I., S. K. Gallagher, M. Hayes, B. A. Lawrence,
plant control strategy in South Africa. Biological et al. 2008. Wetland restoration in the new millen-
Conservation 148: 28–38. (4) nium: do research efforts match opportunities?
Vanbergen and the Insect Pollinator Initiative. 2013. Restoration Ecology 16: 367–372. (10)
Threats to an ecosystem service: pressures on pol- Waterman, C., L. Calcul, J. Beau, W. S. Maet al. 2016. Min-
linators. Frontiers in Ecology and the Environment. 11: iaturized Cultivation of Microbiota for Antimalarial
251–259. doi: 10.1890/120126. (3) Drug Discovery. Medicinal Research Reviews 36(1):
Vandermeer, J., I. Perfecto, and S. Philpott. 2010. Ecologi- 144–168. (3)
cal complexity and pest control in organic coffee Wake, D. B. and V. T. Vredenburg. 2008. Are we in the
production: Uncovering an autonomous ecosystem midst of the sixth mass extinction? A view from
service. BioScience 60: 527–537. (9) the world of amphibians. Proceedings of the National
Varma, V., J. Ratnam, V. Viswanathan, A. M. Osuri, et al. Academy of Sciences USA 105: 11466–11473. (5)
2015. Perceptions of priority issues in the conserva- Waldron, A., R. Justicia, L. Smith, and M. Sanchez. 2012.
tion of biodiversity and ecosystems in India. Biologi- Conservation through chocolate: a win-win for bio-
cal Conservation 187: 201–211. (12) diversity and farmers in Ecuador’s lowland tropics.
Venevsky, S. and I. Venevskaia. 2005. Hierarchical sys- Conservation Letters 5: 213–221. (9)
tematic conservation planning at the national level: Walker, R., E. Arima, J. Messina, B. Soares-Filho, et
Identifying national biodiversity hotspots using al. 2013. Modeling spatial decisions with graph
abiotic factors in Russia. Conservation Biology 124: theory: logging roads and forest fragmentation in
235–251. (7) the Brazilian Amazon. Ecological Applications 23(1):
Venter, O., J. E. M. Watson, E. Meijaard, W. F. Laurance, 239–254. (5)
and H. P. Possingham. 2010. Avoiding unintended Wallace K. J. 2007. Classification of ecosystem services:
outcomes from REDD. Conservation Biology 24: 5–6. (3) problems and solutions. Biological Conservation
Verstraete, M. M., R. J. Scholes, and M. S. Smith. 2009. 139(3–4): 235–246. (3)
Climate and desertification: Looking at an old prob- Waller, D. M. 2015. Genetic rescue: a safe or risky bet?
lem through new lenses. Frontiers in Ecology and the Molecular Ecology 24(11): 2595–2597. (5)
Environment 7: 421–428. (4)
Wallmo, K. and D. K. Lew. 2012. Public willingness to
Vianna, G. M. S., M. G. Meekan, D. J. Pannell, S. P. Marsh, pay for recovering and downlisting threatened and
and J. J. Meeuwig. 2012. Socio-economic value and endangered marine species. Conservation Biology 26:
community benefits from shark-diving tourism in 830–839. (3)
Palau: A sustainable use of reef shark populations.
Biological Conservation 145: 267–277. (3) Wanger, T. C., K. Darras, S. Bumrungsri, T. Tscharntke, et
al. 2014. Bat pest control contributes to food security
Viblanc, V. A., C. Saraux, N. Malosse, and R. Groscolas. in Thailand. Biological Conservation 171: 220–223. (3)
2014. Energetic adjustments in freely breeding–fast-
ing king penguins: does colony density matter? Ward, P. 2004. The father of all mass extinctions. Conser-
Functional Ecology 28(3): 621–631. (6) vation 5: 12–17. (5)
Vidal, O., J. López-García, and E. Rendón-Salinas. 2014. Warkentin, I. G., D. Bickford, N. S. Sodhi, and C. J. A.
Trends in deforestation and forest degradation after Bradshaw. 2009. Eating frogs to extinction. Conserva-
a decade of monitoring in the Monarch Butterfly tion Biology 23: 1056–1059. (4)
Biosphere Reserve in Mexico. Conservation Biology Watanabe, M. E. 2015. The Nagoya Protocol on Access
28: 177–186. (8) and Benefit Sharing: international treaty poses
Vincent, A. C., Y. J. Sadovy de Mitcheson, S. L. Fowler, challenges for biological collections. BioScience doi:
and S. Lieberman. 2014. The role of CITES in the 10.1093/biosci/biv056. (11)
conservation of marine fishes subject to internation- Watson, J. 2013. Endangered species thrive on US mili-
al trade. Fish and Fisheries 15(4): 563–592. (4) tary ranges. Associated Press. August 10. bigstory.
Vranckx, G., H. Kacquemyn, B. Muys, and O. Honnay. ap.org/article/endangered-species-thrive-us-military-
2012. Meta-analysis of susceptibility of woody ranges. (9)
plants to loss of genetic diversity through habitat Watson, J. E., E. S. Darling, O. Venter, M. Maron, et al.
fragmentation. Conservation Biology 26: 228–237. (5) 2016. Bolder science needed now for protected
areas. Conservation Biology. (9)
Bibliography 459

Watson, J. E., N. Dudley, D. B. Segan, and M. Hockings. Wibbels, T. and E. Bevan. 2015. New Riddle in the
2014. The performance and potential of protected Kemp’s Ridley Saga. In State of the World’s Sea
areas. Nature 515: 67–73 (8) Turtles Report. Oceanic Society. (1)
Watson, J. E. M., H. Grantham, K. A. Wilson, and H. P. Wiens, J. J. 2007. Species delimitations: new approaches
Possingham. 2011. Systematic Conservation Plan- for discovering diversity. Systematic Biology 56(6):
ning: Past, Present and Future. In R. Whittaker 875–878. (2)
and R. Ladle (eds.), Conservation Biogeography, pp. Wiersma, Y. F., T. D. Nudds, and D. H. Rivard. 2004. Mod-
136–160. Wiley-Blackwell, Oxford. (6) els to distinguish effects of landscape patterns and
Wearn, O. R., D. C. Reuman, and R. M. Ewers. 2012. Ex- human population pressures associated with species
tinction debt and windows of conservation oppor- loss in Canadian national parks. Landscape Ecology 19:
tunity in the Brazilian Amazon. Science 337(6091): 773–786. (8)
228–232. (5) Wikramanayake, E., E. Dinerstein, J. Seidensticket, S.
Wedekind, C. G. Evanno, T. Székely, M. Pompini, et al. Lumpkin, et al. 2011. A landscape-based conserva-
2013. Persistent unequal sex ratio in a population of tion strategy to double the wild tiger population.
grayling (Salmonidae) and possible role of tempera- Conservation Letters 4: 219–227. (8)
ture increase. Conservation Biology 27: 229–234. (5) Wikström, L., P. Milberg, and K. Bergman. 2008. Monitor-
Weis, J. S. and C. J. Cleveland. 2008. DDT. In C. J. Cleve- ing of butterflies in semi-natural grasslands: Diurnal
land (ed.), Encyclopedia of Earth. Environmental variation and weather effects. Journal of Insect Conser-
Information Coalition, National Council for Science vation 13: 203–211. (6)
and the Environment, Washington, DC. www. Wilcove, D. S. and L. L. Master. 2005. How many endan-
eoearth.org/article/DDT (4) gered species are there in the United States? Fron-
Weiser, E. L., C. E. Grueber, and I. G. Jamieson. 2013. tiers in Ecology and the Environment 3: 414–420. (4, 6)
Simulating retention of rare alleles in small popula- Wilcove, D. S. and M. Wikelski 2008. Going, going, gone:
tions to assess management options for species with Is animal migration disappearing. PLoS Biology 6(7):
different life histories. Conservation Biology 27(2): e188. doi: 10.1371/journal.pbio.0060188. (5, 8)
335–344. (5)
Wildt, D. E., P. Comizzoli, B. Pukazhenthi, and N. Song-
West, P. C., G. R. Narisma, C. C. Barford, C. J. Kucharik, sasen. 2009. Lessons from biodiversity—The value
and J. A. Foley. 2011. An alternative approach for of nontraditional species to advance reproductive
quantifying climate regulation by ecosystems. Fron- science, conservation, and human health. Molecular
tiers in Ecology and the Environment 9: 126–133. (3) Reproduction and Development 77: 397–409. (7)
Western and Central Pacific Fisheries Commission Wilhere, G. F. 2012. Inadvertent advocacy. Conservation
(WCPFC). www.wcpfc.int (6) Biology 26: 39–46. (1)
Western, D., R. Groom, and J. Worden. 2009. The impact Willi, Y., M. van Kleunen, S. Dietrich, and M. Fischer.
of subdivision and sedentarization of pastoral 2007. Genetic rescue persists beyond first-gen-
lands on wildlife in an African savanna ecosystem. eration outbreeding in small populations of a
Biological Conservation 142: 2538–2546. (9) rare plant. Proceedings of the Royal Society B 274:
Whisenant, S. 1999. Repairing damaged wildlands: a process- 2357–2364. (5)
orientated, landscape-scale approach (Vol. 1). Cam- Willis, C. G., B. R. Ruhfel, R. B. Primack, A. J. Miller-
bridge University Press, Cambridge, UK. (10) Rushing, and C. C. Davis. 2008. Phylogenetic
White, P. S. 1996. Spatial and biological scales in reintro- patterns of species loss in Thoreau’s woods are
duction. In D. A. Falk, C. I. Millar, and M. Olwell driven by climate change. Proceedings of the National
(eds.), Restoring Diversity: Strategies for Reintroduc- Academy of Sciences USA 105: 17029–17033. (4)
tion of Endangered Plants, pp. 49–86. Island Press, Wilson, E. O. 1989. Threats to biodiversity. Scientific
Washington, DC. (6) American 261(3): 108–116. (5)
White, T. H., Jr., N. J. Collar, R. J. Moorhouse, V. Sanz, Wilson, E. O. 1992. The Diversity of Life. Harvard Univer-
et al. 2012. Psittacine reintroductions: common sity Press, Cambridge. (10)
denominators of success. Biological Conservation 148:
Wilson, E. O. 2010. Within one cubic foot: Miniature
106–115. (7)
surveys of biodiversity. National Geographic 217(2):
Whitman, W. B., T. Woyke, H. P. Klenk, Y. Zhou, et 62–83. (2)
al. 2015. Genomic Encyclopedia of Bacterial and
Wilson, H. B., E. Meijaard, O. Venter, M. Ancrenaz, and
Archaeal Type Strains, Phase III: the genomes of
H. P. Possingham. 2014. Conservation strategies for
soil and plant-associated and newly described type
orangutans: reintroduction versus habitat preserva-
strains. Standards in Genomic Sciences 10(1): 1. (2)
tion and the benefits of sustainably logged forest.
Whittington, R. J. and R. Chong. 2007. Global trade in PLoS ONE 9(7): e102174. (7)
ornamental fish from an Australian perspective: the
Wilson, M. C., X. Y. Chen, R. T. Corlett, R. K. Didham,
case for revised import risk analysis and manage-
et al. 2016. Habitat fragmentation and biodiversity
ment strategies. Preventive Veterinary Medicine 81(1):
conservation: key findings and future challenges.
92–116. (4)
Landscape Ecology 31(2): 219–227. (4)
460 Bibliography

Winker, K. 2009. Reuniting phenotype and genotype in Wuerthner, G., E. Crist, and T. Butler (eds.) 2015. Protect-
biodiversity research. BioScience 59(8): 657–665. (2) ing the Wild: Parks and Wilderness. The Foundation
With, K. A. 2015. How fast do migratory songbirds have For Conservation. Island Press. (8)
to adapt to keep pace with rapidly changing land- Wunder, S. 2013. When payments for environmental ser-
scapes? Landscape Ecology 30(7) 1351–1361. (4) vices will work for conservation. Conservation Letters
Wofford, J. E., R. E. Gresswell, and M. A. Banks. 2005. 6: 230–237. (9, 11)
Influence of barriers to movement on within-wa- Wünscher, T. and S. Engel. 2012. International payments
tershed genetic variation of coastal cutthroat trout. for biodiversity services: review and evaluation of
Ecological Applications 15(2): 628–637. (5) conservation targeting approaches. Biological Conser-
World Bank. www.worldbank.org (1, 11) vation 152: 222–230. (9)
World Resources Institute (WRI). 2003. World Resources
2002–2004: Decisions for the Earth: Balance, voice, and Xu, H., X. Pan, Y. Song, Y. Huang, M. Sun, and S. Zhu.
power. World Resources Institute, Washington, DC. 2014. Intentionally introduced species: more easily
(11, 12) invited than removed. Biodiversity and Conservation
World Resources Institute (WRI). 2005. World Resources 23(10): 2637–2643. (4)
2005: The Wealth of the Poor—Managing Ecosystems to
Fight Poverty. World Resources Institute, Washing- Yamaoko, K., H. Moriyama, and T. Shigematsu. 1977.
ton, DC. (4). Ecological role of secondary forests in the tradition-
World Wildlife Fund (WWF) Global. wwf.panda.org (6, 9, 11) al farming area in Japan. Bulletin of Tokyo University
World Wildlife Fund (WWF) International. www.world- 20: 373–384. (8)
wildlife.org and www.panda.org (6, 9, 11) Yamaura, Y., T. Kawahara, S. Iida, and K. Ozaki. 2008.
World Wildlife Fund (WWF) and M. McGinley. 2009. Relative importance of the area and shape of
Central American dry forests. In C. J. Cleveland patches to the diversity of multiple taxa. Conserva-
(ed.), Encyclopedia of Earth. Environmental Informa- tion Biology 22: 1513–1522. (8)
tion Coalition, National Council for Science and the Yang, S. L., J. D. Milliman, K. H. Xu, B. Deng, et al. 2014.
Environment, Washington, DC. www.eoearth.org/ Downstream sedimentary and geomorphic impacts
article/Central_American_dry_forests (4) of the Three Gorges Dam on the Yangtze River.
Wright, A. J., D. Veríssimo, K. Pilfold, E. C. M. Parsons, et Earth-Science Reviews 138: 469–486. (4)
al. 2015. Competitive outreach in the 21st century: Yen, S. C., K H. Chen, Y. Wang, and C. P. Wang. 2015.
Why we need conservation marketing. Ocean & Residents’ attitudes toward reintroduced sika deer
Coastal Management 115: 41–48. (12) in Kenting National Park, Taiwan. Wildlife Biology
Wright, G., K. Andersson, C. Gibson, and T. Evans. 2015. 21(4): 220–226. (7)
What incentivizes local forest conservation efforts? Yi, B. L. Islamic clerics issue a fatwa against poachers in
Evidence from Bolivia. International Journal of the Indonesia and Malaysia. Public Radio International.
Commons 9(1). (12) www.pri.org/stories/2016-01-07/islamic-clerics-issue-
Wright, S. 1931. Evolution in Mendelian populations. fatwa-against-poachers-indonesia-and-malaysia (3)
Genetics 16: 97–159. (5) Yochim, M. J. and W. R. Lowry. 2016. Creating conditions
Wright, C. K. and M. C. and Wimberly. 2013. Recent for policy change in national parks: contrasting
land use change inthe western corn belt threatens cases in Yellowstone and Yosemite. Environmental
grasslands and wetlands. PNAS. doi: 10.1073/ Management 1–13. (7)
pnas.1215404110. (4) Yodzis, P. 2001. Trophic levels. In S. A. Levin (ed.), Ency-
Wright, H. L., I. R. Lake, and P. M. Dolman. 2012. clopedia of Biodiversity, Vol. 5, pp. 695–700. Academic
Agriculture—A key element for conservation in the Press, San Diego, CA. (2)
developing world. Conservation Letters 5: 11–19. (9) Young, T. P. 1994. Natural die-offs of large mammals:
Wright, S. J., G. A. Sanchez-Azofeifa, C. Portillo-Quintero, implications for conservation. Conservation Biology
and D. Davies. 2007. Poverty and corruption com- 8: 410–418. (5)
promise tropical forest reserves. Ecological Applica- Young, T. P., T. M. Palmer, and M. E. Gadd. 2005. Compe-
tions 17: 1259–1266. (8) tition and compensation among cattle, zebras, and
Wright, S. J., H. Zeballos, I. Domínguez, M. M. Gallardo, elephants in a semi-arid savanna in Laikipia, Kenya.
et al. 2000. Poachers alter mammal abundance, seed Biological Conservation 112: 251–259. (9)
dispersal, and seed predation in a Neotropical for-
est. Conservation Biology 14: 227–239. (8) Zander, K. K. and S. T. Garnett. 2011. The economic value
Wu, J. and R. J. Hobbs (eds.). 2009. Key Topics in Landscape of environmental services on indigenous-held lands
Ecology. Cambridge University Press, Cambridge, in Australia. PLoS ONE 6(8): e23154. (3)
UK. (8) Zaradic, P. A., O. R. W. Pergams, and P. Kareiva. 2009.
Wu, R., S. Zhang, D. W. Yu, P. Zhao, et al. 2011. Effec- The impact of nature experience on willingness to
tiveness of China’s nature reserves in representing support conservation. PLoS ONE 4: e7367. (11)
ecological diversity. Frontiers in Ecology and the
Environment 9: 383–389. (8)
Bibliography 461

Zellweger, F., V. Braunisch, F. Morsdorf, A. Baltenswei- Zimmermann, N. E., T. C. Edwards, Jr., G. G. Moisen, T.
ler, et al. 2015. Disentangling the effects of climate, S. Frescino, and J. A. Blackard. 2007. Remote sens-
topography, soil and vegetation on stand-scale spe- ing-based predictors improve distribution models
cies richness in temperate forests. Forest Ecology and of rare, early successional and broadleaf tree species
Management 349: 36–44. (2) in Utah. Journal of Applied Ecology 44: 1057–1067. (7)
Zhu, L. and Y. C. Zhao. 2015. A feasibility assessment Zomer, R. J., J. Xu, M. Wang, A. Trabucco, and Z. Li. 2015.
of the application of the Polluter-Pays Principle to Projected impact of climate change on the effec-
ship-source pollution in Hong Kong. Marine Policy tiveness of the existing protected area network for
57: 36–44. (12) biodiversity conservation within Yunnan Province,
Zhu, Y. Y., Y. Y. Wang, H. R. Che, and B. R. Lu. 2003. Con- China. Biological Conservation 184: 335–345. (8)
serving traditional rice varieties through manage- Zydelis, R., B. P. Wallace, E. L. Gilman, and T. B. Werner.
ment for crop diversity. BioScience 53: 158–162. (9) 2009. Conservation of marine megafauna through
Zimmer, C. 2013. Bringing them back to life: The revival minimization of fisheries bycatch. Conservation Biol-
of an extinct species is no longer a fantasy. But is it a ogy 23: 608–616. (4)
good idea? National Geographic 223(14). (7) Zylberberg, M., K. A. Lee, K. C. Klasing, and M. Wikelski.
Zimmermann, A., M. Hatchwell, L. Dickie, and C. D. 2013. Variation with land use of immune function
West (eds.) 2008. Zoos in the 21st Century: Catalysts and prevalence of avian pox in Galapagos finches.
for Conservation. Cambridge University Press, Cam- Conservation Biology 27: 103–112. (4)
bridge. (7)
Index
The letter f after a page number indicates that the entry is included in a figure;
t indicates that the entry is included in a table.

A Allee effect, 186 extinction rates in, 159–160


abiotic environments alleles freshwater, 102–103
invasive species and, 134 definition of, 32 functions of, 72
trophic levels, 34f frequency, 173 invasive species in, 136–137
accidental transport, invasive species natural selection and, 173 ocean acidification, 122–124
and, 133 alpha diversity, 28, 29f plant growth in, 72
acid rain, 117–118, 381 Amazon basin extinctions, 165 restoration of, 353–355
Adams, Ansel, 9–10, 10f Amazon rain forest, edge effects, 111f toxic wastes in, 115
adaptive management, 293, 293f amenity values, 74–77 Arabian oryx (Oryx leucoryx), 239
adaptive restoration, 343–344 American alligator (Alligator missis- arboretums, 256
Addis Ababa Agenda, 383 sippiensis), 224 “arc of deforestation,” 100
Africa, managed communities in, American bald eagle (Haliaeetus Area de Conservación Guanacaste
93–94 leucocephalus), 5, 199f, 224, 313 (ACG), 357, 358f
African elephants (Loxodonta afri- American bison (Bison bison), 266f Argali sheep (Ovis ammon), 197f, 407
cana), 146, 230, 329f American bullfrogs (Lithobates cates- Argentina, ecological footprint of,
African wild dogs (Lycaon pictus), beianus), 134 95f
180, 243 American burying beetle (ABB) arribada, 13
Agency for International Develop- (Nicrophorus americanus), 168f, 370 artificial insemination, 254
ment, U.S., 384 American peregrine falcon (Falco artificial selection, 177–178
“Agenda 21,” 378 peregrinus), 115, 224, 313 Asahiyama Zoo, Japan, 252f
aggregations, vulnerability and, 171 American Samoa, 123f Asian elephants (Elphas maximus),
agricultural conservation, in situ, 325 American Zoo and Aquarium As- 146, 158, 180, 252f
agricultural ecosystems. see also crop sociation, 400f aspirin, 66t
species; farming amphibians assisted colonization, 125, 245
habitat loss and, 97 acid rain and, 117 Association of Zoos and Aquariums,
invasive species and, 133 habitat destruction, 96f 253
land clearance and, 93 maintained in zoos, 250t, 251f Atlantic white cedars (Chamaecyparis
land surface of, 93 Andean mountains, smog in, 55f thyroides), 310
sedimentation and, 117 animals. see also specific animals atropine, 66t
shifting cultivation and, 100 genetically modified species, 142 aurochs (Bos primigenius), 347
subsistence farming, 100 new populations, 238–242 Australasian Regional Association
valuation of, 56f anthropology, methods of, 6 of Zoological Parks and Aquaria,
Aichi Biodiversity Target, 265, antipoaching patrols, 310 253
376–377 Apache trout (Oncorhynchus apache), Australia, rabbits in, 139
air pollution, 55, 117–118, 147 140 Avatar, 404
ajmaline, 66t Apo Island, Philippines, 295f
B
Alaotra grebe (Tachybaptus rufolava- aquaculture, invasive culture and,
background extinction rates, 161
tus), 156 133
backyards “wildlife certification
algae aquariums, 251f, 255–256
program,” 312
blooms, 116f aquatic ecosystems. see also freshwa-
bacteria, species of, 44
cultural eutrophication, 353 ter habitats
symbiotic, 36 artificial ponds as, 292
464 Index

bald eagle (Haliaeetus leucocephalus), biodiversity conservation models, bowhead whales (Balaena mysticetus),
310, 313 287t 131t
baleen whales, 131t biogeography model, 278 Braulio Carrillo National Park, Costa
banded kingfisher (Lacedo pulchella), bioimagnification, 113 Rica, 285
165f Biological Abstracts, 195 Brazil, reverence for nature, 11
Banff National Park, Canada, 284 biological communities, 33 Brazilian Amazon
Banhine National Park, 382f managed, 93–94 deforestation and, 100
Barksdale Air Force Base, Louisiana, species diversity and, 16 indigenous lands, 323f
310 biological control, invasive species Brazilian free-tailed bat (Tadarida
barred tiger salamander (Ambystoma and, 133 brasiliensis), 406, 406f
mavortium), 27 biological diversity. see biodiversity Brazilian ocelot, 254f
Barro Colorado Island, Panama, 158f Biological Dynamics of Forest Frag- British Broadcasting Corporation,
Bat Conservation International mentation (BDFF) study, 107f 404
(BCI), 405 biological field stations, 267 brown bats (Myotis lucifugus), 144f
bats biological products, trade in, 227, 374 brown pelican (Pelecanus occidenta-
importance of, 406 biological species, 25 lis), 81, 224
white-nose syndrome, 143–144 biomes, human impacts on, 98f brown tree snake (Boiga irregularis),
bees, 67, 73f biopiracy, 79 136
behavior, population biology and, bioprospecting, 78 Buddhism
195 bioregional management, 320 conservation biology and, 7
behavioral ecology, 242–243 bioremediation, 348 stewardship responsibilities, 85
beluga whale, 131t BIOSIS Citations Index, 195 buffalo commons, 356
beneficiary values, 80–83 biosphere reserves, 296–298 Buijs, Tina, 284f
bequest values, 80–83 biota, definition of, 33–34 Bureau of Land Management, U. S.,
Bern Convention, 222 biotic barriers, 343f 317, 372
beta diversity, 29, 29f biotic interactions, 194 Bush Blitz, 43
Bhutan Trust Fund for Environmen- bird monitoring online, 394 bushmeat, 63, 128
tal Conservation, 386–387 Birdlife International, 220 butterflies
bighorn sheep (Ovis canadensis), 206, birds abundance of, 158
209–210 extinction rates, 157f impacts of weather on, 184
binomial nomenclature, 26 habitat destruction, 96f monarch, 395f
bioblitz, 43 maintained in zoos, 250t overharvesting of, 128
biocontrol, 348 trade in, 129t C
biocultural restoration, 357–358 birth rates, demographic variation
cacti, trade in, 65, 128, 129t
biodiversity and, 185
caffeine, 66t
aesthetic benefits of, 86 black-footed ferrets (Mustela ni-
California condors (Gymnogyps
benefits of, 227–228 gripes), 142–143, 238, 239, 240f
californianus), 238, 243f
efforts, 339 black howler monkeys, 327
California tiger salamander (Ambys-
environmental ethics of, 87 black-legged ticks, 110, 145f
toma californiense), 27, 28f
ethical values of, 83–86 black rhino (Diceres bicornis), 304f,
Cambrian Period extinctions, 152f
indicators, 29f, 219 333
Canada, ecological footprint of, 95f
indirect use values, 87 blue whales (Balaenoptera musculus),
Candidatus xenohalitosis californiensis,
information availability, 395–396 131
160f
international benefits, 227–228, bluefin tuna (Thunnus thynnus), 126f
canine distemper virus, 142–143
374 body size, home ranges and, 169
Cape Cod heathland, Massachusetts,
international funding, 384f bohemian waxwing (Bombycilla gar-
291f
laws protecting, 366–367 rulus), 36f
Cape Floristic Region, South Africa,
levels of, 23, 24f Border areas, 312
46f
losses of, 3, 92f Borneo, forest destruction in, 102
captive breeding programs, 238–239,
option values of, 87 Bosnia and Herzegovina, protected
249–250
protection of, 394–396 areas, 270
carbon cycle, human impacts on, 93t
recreational benefits of, 86 Botanic Gardens Conservation Inter-
carbon dioxide (CO2)
threats to, 90–146 national (BGCI), 258
anthropogenic, 123
values, 52–87 botanical gardens, 256–258
atmospheric concentrations, 120f
worldwide, 41–48
Index 465

climate change and, 118–119 co-management, 322 local regulations, 365–367


global ecosystems and, 226 coastal development, coral reefs and, measures, 371–372
plant absorption of, 72 105 national level, 372–373
Carboniferous Period extinctions, coca (Erythrozylum coca), 66t outside protected areas, 304–335
152f cocaine, 66t prioritization, 216–221
carnivores, trophic levels, 36 codeine, 66t public benefits, 371–372
Carolina anoles (Anolis carolinensis), coffee plantations, 316, 316f, 397f regional regulations, 365–367
109 colonization species approach, 217
carrying capacity (K), 35, 184 assisted, 125, 245 wilderness approach, 218–219
Carson, Rachel, 9, 9f, 113 European, 132–133 Conservation, Amenity and Recre-
catastrophes, environmental varia- extinction rates and, 163f ation Trusts (CARTS), 367
tion and, 187–188 limits to, 107–108 conservation banking, 369–370
cats, domesticated, 109, 134 Columbia Basin pigmy rabbit conservation biologists
Caulerpa taxifolia, 137 (Brachylagus idahoensis), 248, 248f challenges for, 402–403
“Cecil the Lion,” 53, 298, 327 commercial development, endan- as managers, 408
censuses. see also surveys gered species and, 97 as motivators, 407–408
description of, 200–201 commodity values, 60 role of, 402–408, 403f
population monitoring with, 199 common-property resources, 127. see conservation biology
cerulean warbler (Dendroica cerulea), also open-access resources basic sciences and, 7f
109 regulation of, 55 definition of, 2–18
cetaceans, conservation of, 256 common witch hazel (Hamamelis ethical principles, 15–17
charismatic megafauna, 80–81, 252 virginiana), 65 future of, 17–18
cheatgrass (Bromus tectorum), 134 Commons, Open Spaces and Foot- goals of, 5
checkered-skipper butterfly (Pyrgus paths Preservation Society, 10 roots of, 6–12
oileus), 158f communities, compositions of, 35–36 science of, 5–15
cheetah (Acinomyx jubatus), 333 Community Baboon Sanctuary, conservation concessions, 371
Cheonggyecheon Stream, Seoul, Belize, 327 conservation corridors, 284
Korea, 345f community-based conservation conservation covenants, 368
Chesapeake Bay, clean-up projects, (CBC), 327 conservation development, 368–369
353–355, 354f Community-Based Natural Resource conservation easements, 367–368,
Chicago Wilderness project, 283 Management (CBNRM), 326, 369f
Chile, 55f, 284f 330–333 conservation leasing, 369
chimpanzees, studies of, 310 community conserved areas, 327 conservation organizations, funding
China compensatory migration, 339 for, 386f
coral reef destruction in, 104–105 competition, species interactions, 35 conservation planning units, 377f
ecological footprint of, 95f cone snails, 67 conservation psychology, 367
managed biological communities Conference on Environment and conservation strategies, ex situ,
in, 93–94 Development, UN, 375 246–261
chinchilla (Chinchilla spp.), 128 Conference on Sustainable Develop- Consultative Group on International
Chinese giant salamander, 251f ment, UN, 375, 378 Agricultural Research (CGIAR),
Chitwan National Park, Nepal, 183 Conference on the Human Environ- 258, 325
Christianity, conservation and, 7 ment, UN, 374 consumptive use values, 61–63, 62f
cichlid fishes, 136 coniferous forests, case studies, 109 Convention Concerning the Protec-
cinerous vultures (Aegypius mona- conservancy tion of the World Cultural and
chus), 407 benefits of, 332 Natural Heritage, 379
citizen-science programs, 394 ongoing problems, 394–402 Convention for the Conservation
Civilian Conservation Corps, 356f conservation and Management of Highly
Clean Water Act, 351 categories, 212–215, 213f Migratory Fish Stocks in the
clear-cutting, edge effects and, 109 ecosystem approach, 218 Western and Central Pacific
climate change, 299–300 enforcement, 371–372 Ocean, 231
extreme weather events, 122 financing of, 59–60 Convention on Biological Diversity
global, 118–126 funding effectiveness, 383–389 (CBD), 12, 57, 260, 375f, 376–379,
Internet monitoring, 394 global scope of, 396 379t, 396
climate regulation, plants in, 72–73 hotspot approach, 219–221
climatologists, 6, 119–120 local level, 399–400
466 Index

Convention on International Trade D limits to, 107–108, 169–170


in Endangered Species (CITES), Darwin, Charles, 155, 166 seed, 108, 344
128, 228–231, 375, 375f, 379t, 396 data deficient (DD) category, 213 distinctiveness, 217
enforcement of, 222 de-extinction programs, 261–262 distribution, population biology
Convention on Long-range Trans- dead zones, 116 and, 194
boundary Air Pollution, 381 death rates, demographic variation disturbances, human contact and,
Convention on the Conservation and, 185 171
of Antarctic Marine Living Re- debt-for-nature swaps, 387 Djoudj National Bird Sanctuary, 380f
sources, 231 decomposers, trophic levels, 34f, 37, DNA analysis
Convention on the Conservation 344 bacterial, 44
of Migratory Species of Wild deep ecology, 86–87 environmental, 201–202
Animals (CMS), 231, 379t Deepwater Horizon oil spill, 15, 394 extinct species, 261
Convention on the Law of the Sea, deer (Odocoileus spp.), 35, 39 population surveys and, 201–202
381, 382–383 Defenders of Wildlife, 250 dogs, domestic (Canis lupus familia-
Convention on the Prevention of deforestation ris), 136
Marine Pollution by Dumping of arc of, 100 domesticated animals
Wastes and Other Matter, 383 extinctions related to, 158 diseases in, 109–110
Convention on Wetlands, 379t rates of, 93 medicines used for, 115
Cope’s gray tree frog (Hyla chrysoce- degazettement, 270 outbreeding depression and,
lis), 27f degradation, 299 177–178
Coquerel’s sifaka (Propithecus co- demilitarized zones, 312 rare breeds, 250
quereli), 169f demographic stochasticity, 184–187, dragonflies, habitats for, 314
coral reefs 205 drilling projects, 44
bleaching, 123, 123f demographic studies, population drinking water contamination, 70
temperature limits, 123 surveys, 199, 203 drugs, natural pharmacy, 66–67, 66t
threats to, 104–105, 104f demographic variation, 185 Ducks Unlimited, 81
Cornell Lab of Ornithology, 394 demography, population biology DUE acronym, 217
cost-benefit analyses, 57 and, 195 dusky seaside sparrow (Ammodra-
cost-effectiveness analyses, 59 Denmark, protected areas, 270 mus maritimus nigrescens), 186
Costa Rica Denver, John, 10 E
Merck contract, 260 desert tortoises (Gopherus agassizil), Earth Summit, Rio de Janerio, 12, 376
Ministry of the Environment, 357 310 Ebola virus, 146
reverence for nature, 11 desertification, 105–106 ecocolonialism, 322
tropical rain forest restoration, detritivores, trophic levels, 37 ecological economics, 56
357–359 developed countries ecological footprints
wildlife reserves, 285 conservation projects, 396–397 definition of, 94–95
cowbirds, nest parasitization of, 109 park management and, 294 of nations, 95f
coyotes (Canis latrans), 26 support from, 397 Ecological Management and Restora-
Cretaceous Period extinctions, 152f development, views of, 95–96 tion, 359
critically endangered (CR) category, development projects, evaluation of, ecological resilience, 337
213 397–398 ecological restoration, 337
crop species, 325. see also agricultural Devonian Period extinctions, 152f ecological succession, 289
ecosystems diatoms, 45f ecologically extinct, 155
climate change and, 125 dichloro-diphenyl-trichloro-ethane economic development
genetic diversity in, 260f (DDT), 113 description of, 363–365
genetically modified, 142 digitoxin, 66t planning processes, 403
landraces, 259 dinoflagellates, 45f sustainable, 364
species relationships and, 73–74 direct use values, 60, 61, 82f economic growth, 363
cross-fostering, 254 disease transmission economic use values, 61–78
culling invasive species, 140–141 human activities and, 142–146 economic values, wetlands, 82f
cultural eutrophication, 353 management of, 144–145 economics
cultural geography, role of, 6 movement corridors and, 286 definition of, 54
cultural services, 61 diseases, introduction of, 128 ecological, 55–61
cyclosporine, 79 dispersal environmental, 55–61
highway overpasses, 285f
Index 467

ecosystem approach, 218 endangered (EN) species, 213, 224f, European Union Biodiversity Con-
ecosystem diversity, 33–41 225f vention, 339
basis of, 24–25 human activities and, 97 eutrophication, 116, 353
description of, 23, 24f isolated populations of, 175–176 Everglades Headwaters National
understanding of, 16 legal definition, 223 Wildlife Refuge, 368f
ecosystem dynamics, 41 reintroduction programs, 241 evolutionary flexibility, loss of,
ecosystem engineers, 39 Endangered Species Act (ESA), U. S., 178–179
ecosystem function, restoration 201, 317 evolutionary species, 25
thresholds, 343f experimental populations, ex situ conservation strategies,
ecosystem health, 400–401 245–246 246–261, 247f
ecosystem management, 318–321 function of, 222–226 existence values, 60, 80–83, 82f,
definition of, 9 on private forested lands, 308 227–228
themes in, 318, 319f recovery criteria, 206 experimental essential populations,
ecosystem services, 60–61, 73f endangerment, 218 245
ecosystems endemic species, 47, 164–165 experimental nonessential popula-
agricultural, 38f energy capture, photosynthesis and, tions, 245
biological, 3 33, 34f extant species, 155–157
definition, 33 energy efficiency, 365 externalities, definition of, 54
healthy, 41 energy transfer, trophic levels and, extinct, meaning of, 154–160
prairie, 70f 36 extinct (EX) category, 213, 213f
preservation of, 266 Environment Canada, 200 extinct in the wild (EW) category,
productivity of, 68–70 Environmental Defense Funds, 386f 154, 213
stable, 41 environmental economics, 56 extinct species
structure and functions of, 341f environmental ethics, 6, 83–87 near relatives of, 171
ecotourism environmental justice, 85 technology in return of, 261–262
dangers of, 77 environmental monitors, 73–74 extinction debt, 157
description of, 75 Environmental Performance Index extinction rates
economic justification of, 76f (EPI), 58, 59f habitat loss and, 164–166
Eden Project, England, 256 environmental stochasticity, 184, predictions, 161–164
edge effects 187–188, 205 trends in, 155–156, 157f
case studies, 109 environmental trust funds, 386–387 extinction vortices, 188–189, 189f
definition of, 106 environmental variation, extinction extinctions, 150–191
habitat fragmentation and, and, 188 cascades, 39–40
110–112 environmentalism, 6 climate change and, 125
interspecies interactions and, environments habitat fragmentation and,
109–110 carrying capacity (K), 184 108–109
protected area, 283 forms of damage to, 112–118 local, 157–159
educational values, 77 population biology and, 194 measurement of, 160–166
educators, conservationists as, 403 ephedrine, 66t prevention of, 15, 166
effective population size (Ne), 174. Equator Initiative of the United Na- vulnerability to, 166–172
see also population sizes tions, 330 extirpated species, 154
determination of, 179–184 erosion, flooding and, 70 extractive reserves, 325–327, 326f
loss of genetic variation and, 175 essential minerals, levels of, 116 F
Ehrenfeld, David, 171–172 establishment programs, 241,
Facebook, 405
Ehrlich, Paul, 12 243–244
Fair Trade Certified standard, 398
elephant seals, sex ratios, 180 Etosha National Park, southern
farming. see also agricultural ecosys-
elephants, 299f, 310. see also African Africa, 212f
tems
elephants (Loxodonta africana); Eurasian thistles (Carduus spp.), 133
habitat management and, 56f
Asian elephants (Elphas maximus) Europe. see specific countries
subsistence-level, 100
elk (Cervus elaphus), 180, 180f European bison (Bison bonasus),
farmland restoration, 355–357
embryo transfer, 254 242–243
Fauna Europaea database, 222
Emerson, Ralph Waldo, 8 European colonization, 132–133
fencing, in restoration ecology, 338f
“empty forests,” 40 European honeybees (Apis mellifera),
field biology, 372f
Encyclopedia of Life project, 44, 396 135
field ecosystems, 34f
468 Index

fieldwork, population studies, four Rs, 277–278 loss of, 170


197–198 fragmentation effects population sizes and, 171–176
filmmakers, 404 climate and, 112 genetic drift
fire, in protected areas, 289 protected area size and, 283–286 description of, 173
fire ants (Solenopsis invicta), 135, 135f Francis, Pope, 18 habitat fragmentation and, 109
Fish and Wildlife Service (FWS), U. Franklin Tree (Franklinia alatamaha), loss of genetic variation and, 173f
S., 223, 369–370, 384, 394 247 genetic swamping, 140
fishes Fresh Kills landfill, New York, 346 genetic variation
extinctions, 159 freshwater habitats. see also aquatic management of, 183–184
pollution threats to, 115 ecosystems species decline and, 170
fishing extinctions in, 159–160 genetically modified organisms
commercial, 130–132 threats to, 102–103 (GMOs), 141–142
dietary protein and, 63 Frodo effect, 282 genetics, population biology and,
exclusion zones, 198 frogs, fungal diseases, 142 195
flagship species, 217, 317 Frogwatch, USA, 200 genome resource bank (GRB), 254
flamingos (Phoenicopterus ruber, P. frontier forests, loss of, 97 genotypes, 32
minor), 211 fuelwood, consumptive use value, 63 Genuine Progress Indicators (GPI),
flooding, impacts of, 71 functional diversity, 28 58
Florida panther (Felis [Puma] concolor functionally extinct, 155 geographic information systems
coryi), 308, 309f fundamental theorem of natural (GIS), 276–277, 276f
focal species, 217 selection, 179 geographic ranges, 167
food, restricted access to, 108 funding geography
Food and Agriculture Organization for conservation areas, 383–389 cultural, 6
of the United Nations (FAO), 226 for conservation organizations, physical, 6
food chains, 37–38, 114f 386f German Federal Agency for Nature
food webs, 37–38, 38f effectiveness of, 387–389 Conservancy, 310
Forest Alliance program, 385 long-term, 388 Germany
Forest Carbon Partnership Facility, political activity and, 406 habitat modification in, 97
385 short-term, 388 protected areas, 270
forest products industry, 65 fungal diseases, 142–144 giant panda (Ailuropoda melanoleuca),
Forest Stewardship Council, 317, 184–185, 253f, 329
G
388, 398 giant tortoise (Chelonoidis elephanto-
Galápagos Islands, Ecuador, 155
forestry, valuation of, 56f pus), 155
gamma diversity, 28, 29f
forests. see also specific forest types Glacier National Park, U. S., 382
Gandhi, Mahatma, 85
“arc of deforestation,” 100 glaciers, melting of, 123, 124t
gap analyses, 274–277
carbon retention by, 72–73 Glanville fritillary butterfly (Melitea
garlic mustard (Alliaria petiolata), 188
case studies, 109 cinxia), 175f
gastric-breeding frog (Rheobatrachus
destruction of, 102 GlaxoSmithKline corporation, 78–79
silus), 261
edge effects, 111f Global Biodiversity Information
geese, sex ratios, 180
“empty forests,” 40 Facility, 44
gene flow
forest products industry, 65 global climate change, 118–126
inbreeding and, 176
fragmented, 110 global ecosystems, human impacts
between populations, 174–175
frontier, 97 on, 93t
General Assembly Special Session on
logging and, 101 Global Environment Facility (GEF),
the Environment, 375
on private lands, 308 384, 384f
genes (loci)
protected, 272–273f global warming, effects of, 124–126,
diversity and, 31
restoration, 357–359 124t
mutation rates of, 174
species losses, 165 globalization, interconnection and,
genetic diversity, 31–33
unprotected, 310 94
benefits of, 24
Formosan sika deer (Cervus nippon globally extinct, 154
consequences of reduction in,
talouanus), 246f glyphosate, 142
176–179
Fort McCoy, Wisconsin, 310 Go Diego Go, 404
crop species, 260f
fossil fuels, human use of, 119 Gold King Mine, 113f
description of, 23, 24f
fossil records, 161 golden-cheeked warbler (Dendroica
isolated populations and, 175–176
founder effects, 182 chrysoparia), 177f
Index 469

golden lion tamarins (Leontopithecus habitat destruction, 96–106 horticulture, invasive species and,
rosalia), 239, 243 climate change and, 125 133
Gonarezhou National Park, 382f disease transmission and, 144–145 hotspot approach, 219–221
Google Scholar, 196 habitat disturbances hotspots, 220, 221f
Gordon and Betty Moore Founda- impacts of, 97 Hudson River School, 9
tion, 385 tolerance for, 171 human activities. see also humans
gorillas, studies of, 310 habitat fragmentation, 106–112, 372f altered landscapes and, 307f
grasslands description of, 97, 106–107 CO2 concentrations and, 120f
grazing, 69f species persistence, 166 disease transmission and, 142–146
species diversity and, 288–289 threats posed by, 107–110 mass extinction related to, 155–157
threats to, 102 habitat islands, 278 sixth extinction episode and, 153
unprotected, 310–311 habitat loss species interactions with, 195
gravel pits, 314 description of, 97 threats to biodiversity from, 92f
gray tree frog (Hyla versicolor), 27f edge effects, 110f West Africa, 311f
gray whale (Eschrichtius robustus), extinction rates and, 164–166 human-animal conflicts, 299
131, 131t fragmentation and, 110f human diversity, 85
gray wolf (Canis lupus), 26, 39, 236, IUCN categories and, 214–215 human-dominated landscapes
237f, 238 habitat/species management areas, biodiversity in, 306
grayling fish (Thymallus thymallus), 268t conservation in, 313–318
181 habitat steppingstones, 306f human immunodeficiency virus
Great Barrier Reef Marine Park, 75 habitat/supporting services, 61 (HIV), 146
Great Lakes, North America, 137 habitats, 35 human life, respect for, 85
Great Smoky Mountains National unprotected, 307–313 Human Microbiome Project, 44
Park, U. S., 380 hair samples, DNA analysis of, human population
great white pelicans (Pelecanus ono- 201–202 formula for impact of, 94
crotalus), 380f Haiti, protected areas, 268 impacts of growth in, 92–96
Greater Limpopo Transfrontier Park, Hallwachs, Winnie, 357 inbreeding depression in, 176
382, 382f hard release programs, 240–241 increases in, 3, 4f
greater one-horned rhinoceros (Rhi- Hardin, Garett, 54–55 wealthy country, 59f
noceros unicornis), 183 hatching failures, 177f humanitarian organizations, 398
Green Island, Kure Atoll, 200f Hawaiian monk seal (Monachus humans, dominance of, 155
“green products,” marketing of, 388 schauinstandi), 200, 200f humpback whale (Megaptera novae-
“green technology,” 365 Hawaiian stilt (Himantopus mexicanus angliae), 131t
greenhouse effects, 119 knudseni), 204 Hungary, protected areas, 270
greenhouse gas emissions, 119 head-starting populations, 238 hunting
Paris Agreement on, 13f heat waves, 124t conservation and, 60
Greenland, protected area, 269 herbivores, trophic levels, 35, 36 dietary protein and, 63
“greenwashing,” 365 hermaphroditic species, 185–186 intensive, 127–128
grey partridge (Perdix perdix), 202f herpetofauna, 348–349 species declines and, 169
grizzly bears (Ursus arctos horribilis), heterozygosity, 32–33 Hurricane Katrina, 71, 351
227 population size and, 173–174 Hwange National Park, Zimbabwe,
gross domestic product (GDP), risk of extinction and, 175f 53
57–58, 59f Hinduism hybrid vigor, 178
Guam, forest bird species, 136 conservation biology and, 7 hybridization, 25–26, 27
Guanacaste Dry Forest Conservation stewardship responsibilities, 85 hybrids, production of, 27
Fund, 357 HIV/AIDS, trends in, 78–79 hydrocarbons, 118
guilds, competing species, 37–38 home gardens, 312 hyenas, sex ratios, 180
gymnosperms, habitat destruction, homeowners associations, 312–313, I
96f 313f
ibex (Capra ibex), 178
gypsy moths (Lymantria dispar), 135 Homo sapiens. see humans
Iguaçu Falls, Brazil, 380
H homozygous populations, 33
Ikh Nart Nature Reserve, Mongolia,
horseshoe crabs (Limulus polyphe-
habitat conservation plans (HCPs), 407
mus), 77–78, 78f
226 Important Bird Areas (IBAs), 220
habitat corridors, 284–286
470 Index

in situ conservation strategies, International Union for Conserva- Jurassic Period extinctions, 152f
246–261 tion of Nature (IUCN), 212–215, K
iNaturalist, 394 213f, 226, 388, 404
Kakamega Environmental Education
inbreeding classification system, 267
Program (KEEP), 8f
costs of, 176 Plant Conservation, 220
Kalimantan, Indonesia, deforestation
gene flow and, 176 Red List, 213, 213t
and, 100
inbreeding depression, 176–177 International Whaling Commission
Kamchatka brown bear (Ursus arctos
hatching failures and, 177f (IWC), 130–131
beringianus), 187f
population sizes and, 108–109 International Year of Forests, 12
Karner blue butterfly, 310
Index of Sustainable Economic Wel- International Zoo Yearbook (IZY), 249
Kasigu Corridor Reducing Emissions
fare (ISEW), 58 Internet monitoring, 394
from Deforestation and Forest
India Internet resources, population stud-
Degradation (REDD), 329
ecological footprint of, 95f ies, 195–196
Kemp’s ridley sea turtle (Lepidochelys
managed biological communities interspecies interactions, 109–110
kempii), 13, 14f
in, 93–94 intrinsic values
Kenya, wildlife conservation,
resource consumption in, 94 of biodiversity, 227–228
330–333
Indiana Classified Forest and Wild- biological diversity, 15
keystone resources, 40–41, 292
lands Program, 312 definition of, 84
keystone species
indicator species, 217 introduction programs, 236
parks, 281
indigenous lands, 322–323, 323f invasive species, 132–142
protection of, 38–41
indigenous people. see traditional in aquatic habitats, 136–137
killer whales (Orcinus orca), 404
people characteristics of, 137–140
king penguins (Aptenodytes patagoni-
indirect use values, 60, 68, 82f, 87 control of, 140–141
cus), 197–198
Indonesia, reverence for nature, 11 hybridization of, 140
Kirstenbosch National Botanical
Industrial Revolution, 92 on oceanic islands, 135–136
Garden, Cape Town, 257f
industrialized countries, resource threats posed by, 134–135
kiwi (Apteryx spp.), 136
use by, 94 ipecac, 66t
Komodo dragon (Varanus komodoen-
infectious disease transmission, 143f Iraq
sis), 217f, 380
infrastructure projects, endangered Landsat images, 352f
Kruger National Park, 382f
species and, 97 protected areas, 268
Kuna people, Panama, 297f, 324f
insects, niche variability, 35–36 Islam, conservation beliefs, 7
Kuna Yala Indigenous Reserve,
integrated conservation and devel- island biogeography model, 278,
Panama, 297, 324
opment projects (ICDPs), 322, 330 279f, 280
Kunming Botanical Garden, 258
integrated pest management (IPM), colonization and extinction rates,
Kure Atoll, 200–201
141 163f
Kyoto Protocol, 1997, 378
Inter-American Development Bank, extinction rates, 161–164
385 species counts, 162f L
Intergovernmental Panel on Climate island habitats, extinction and, La Selva Biological Station, Costa
Change (IPCC), 119 167–168 Rica, 285
international agreements, 226–231, ivory-billed woodpecker (Campephi- Lake Erie restoration, 353
379t lus principalis), 156 Lake Michigan, Great Lakes, 137f
International Biodiversity Day, 12 ivory trade, 230, 230f Lake Titicaca frog (Telmatobius co-
International Conference on Conser- leus), 129, 129f, 400f
J
vation Biology, 1978, 12, 405f Lake Victoria, East Africa, 136
Janzen, Daniel, 357, 358f
International Conference on Financ- Lake Worth, Florida, 313f
Japan
ing for Development (FFD3), 383 land, undesirable to humans, 312
ecological footprint of, 95f
International Convention for the Land and Water Conservation Fund,
reconciliation ecology, 345
Regulation of Whaling, 231 372
Japanese crested ibis (Nipponia nip-
International Crane Foundation, 253 land ethic, description of, 9
pon), 177f
International Species Information land sharing, 315
jellyfishes, 139
System (ISIS), 253 land sparing, 315
The Journal of Ecosystem Restoration,
international trade in biologicals, 227 land surface, human impacts on, 93t
359
International Treaty on Plant Genetic land trusts, 367–371, 368f
Journey North website, 395f
Resources for Food and Agricul- landfills, restoration projects, 346
Judaism, 7, 85
ture, 379t landowners
Index 471

conservation easements and, 368 The Lord of the Rings, 404 migrations. see also specific Conven-
conservation leasing, 369 Louisiana Coastal Wetlands Con- tions
deforestation and, 100–101 servation and Restoration Task across borders, 227, 374
legal titles, 322 Force, 351 compensatory, 339
landraces, 259 Lubchenco, Jane, 402 disease transmission and, 144
Lands Legacy Initiative, 372 Lyme disease, 110, 145f, 315 elephant, 329f
Landsat images, Iraq, 352f lynx (Lynx canadensis), 227 movement corridors and, 284–285
landscape design, 287t between protected areas, 305
M
landscape ecology, 286–288 seasonal, 170
Magellanic penguins (Spheniscus
landscapes songbird, 109
magellanicus), 198, 198f
human-altered, 307f military lands, biodiversity on,
magnolia tree (Magnolia sinica), 257
human-dominated, 315–318 308–310
Malaysia, deforestation and, 100
landslides, 70 Millennium Ecosystem Assessment,
Malpai Borderlands Group, 319
Larinus planus (herbivorous weevil), 61
mammals
133 Millennium Seed Bank, Kew, 257f
extinction rates, 280f
Las Condes, Santiago de Chile, 55f minimum dynamic area (MVA), 208
habitat destruction, 96f
leaders minimum viable populations
maintained in zoos, 250t
conservationists as, 403 (MVPs), 206–208, 207f
native species, 48t
politically active, 406 minke whales, 131t
managed resource protected areas,
least Bell’s vireo (Vireo bellii pusillus), Missouri Botanical Garden, Mada-
268t
310 gascar program, 257f
management plans, 275
least concern (LC) category, 213 mitigation activities, 246
managers, collaboration among, 17
legal titles, 322 molluscs, overharvesting of, 128
manatee (Trichechus manatus), 223,
legislation monarch butterflies, 395f
256
future of, 400–401 Monteverde golden toad (Bufo peri-
mangrove forests, 103–104
international, 226–231 glenes), 156
Manu National Park, Peru, 380
national, 222–231 morphine, 66t
marine coastal waters, 103
lemurs, 169f morphological species, 25
marine protected areas (MPAs), 203,
Leopold, Aldo, 80 morphology, information about, 194
270f, 271–272, 271f, 294–296
lesser adjutant (Leptoptilos javanicus), morphospecies, 25
marine restoration projects, 355
370f mosquitoes, control of, 113
market failures, 55
leukemia, drugs for, 67 motivators, conservationists as, 403
mass extinctions, 151
lidocaine, 66 mountain ash tree (Sorbus aucuparia),
mates, access to, 108
limited development, 368–369 36f
Mauritius kestrel (Falco punctatus),
limiting resources, 35 movement corridors, 284
183, 240
Limpopo National Park, 382f Muddy Boggy Conservation Bank,
maximum sustainable yields, 130
limulus amoebocyte lysate (LAL), 78 Oklahoma, 370
McNeely, Jeff, 60, 306
lions (Panthera leo), 182 Muir, John, 8
Mead’s milkweed (Asclepias meadii),
livestock grazing, 97 Muir Woods National Monument,
244
Living Planet Index, 215 291f
measurement units, 11t
local-is-best (LIB) approach, 341–342 mules, 177–178
Mediterranean Action Plan (MAP),
local peoples, working with, multinational environmental agree-
320, 320f
321–330, 399–400 ments (MEAs), 375f
Mediterranean-type communities, 46
locally extinct, 154 multiple use habitats, 317
megafauna, definition of, 155
loggerhead shrike (Lanius ludovicia- Munich Botanical Gardens, Ger-
metapopulations
nus anthonyi), 208f many, 257f
description of, 107, 208–210
logging mutations, genetic variation and,
population sizes and, 210f
climate change and, 119 32, 174
viability analysis, 245f
endangered species and, 97 mutualism, 35
methane, climate change and,
rain forest loss and, 101
118–119 N
restrictions on, 373
Mexico, sea turtle recovery efforts, 14 Namibia, wildlife conservation, 59,
sedimentation and, 117
microclimates, edge effects and, 330–333, 331f
selective, 310
111–112 narwhale, 131t
Long-Term Ecological Research
National Audubon Society, 367, 394
(LTER), 210–212, 211f, 293
472 Index

National Cancer Institute, U.S., 78 NGO Conservation International, orchids


National Climate Action Plans, 378 385–386 habitat requirements, 171
National Conservation Areas, 372 Ngorongoro Crater lions, 182, 183f trade in, 128, 129t
national environmental funds niche requirements, 170–171 Ordovician Period extinctions, 152f
(NEFs), 386 Nile perch (Lates niloticus), 136 organochlorine pesticides, 113
National Forest Conservation Pro- nitric acid, 117–118 ornamental fish, trade in, 129t
gram, China, 329 nitrogen cascade, 118 otters, impacts of, 242
National Forest Management Act, nitrogen cycle, human impacts on, outbreeding depression, 177–178,
317 93t 341–342
national forests, protected, 272–273f nitrogen deposition, 118 overexploitation, 126–132
National Marine Fisheries Service nitrogen oxides, 117–118 overfishing, 57, 126f
(NMFS), 223 “no net loss policies,” 339 overgrazing, 69
national monuments, 268t, 372 non-use values, 60 overharvesting
National Park Service, U. S., 140–141, non-wood products, forest, 65 risk of extinction and, 171–172
372 nonconsumptive uses, wildlife, 68 species declines and, 169
national parks nongovernmental organizations overseas development assistance
description of, 268t (NGOs), 383–386 (ODA), 383–386
picnic areas, 74f local, 401 Ozone Layer, 381
National Science Foundation, U.S., nonpoint source pollution, 353 ozone production, 118
293, 384 normative disciplines, 5 P
National Trust, protected lands, 367 North American Amphibian Moni-
Pacific Remote Islands Monument,
National Trust for Places of Historic toring Program, 200
272
Interest or Natural Beauty, 11 North American beaver (Castor
Pacific salmon, captive populations,
National Wildlife Federation, 312 canadensis), 40f
239
National Wildlife Refuges, 372 North American Bird Survey (BSS),
Padre Island National Seashore,
Native Americans, conservation 202f
Texas, 14–15
beliefs, 7 North American Breeding Bird
Panama Amphibian Rescue and
Native Seeds/SEARCH, 325 Survey, 201
Conservation, 250
native species, reestablishment of, North American thistles (Cirsium
“paper parks,” 100, 288
339 spp.), 133
Papua, New Guinea Reserves, 322
natural catastrophes, 187–188 North Korea demilitarized zone, 312
parasites, epidermiology of, 37
Natural Heritage programs, 158 northern bobwhite (Colinas virginia-
parasitism, invasive species and, 134
natural history, 194 nus), 135f
Paris Accord, 378
natural mass extinctions, 151 northern right whale, 131t
Paris Agreement, 12
natural resources not applicable (NA) category, 214
Paris Climate Conference, 375
impact of human populations not evaluated (NE) category, 214
park management
on, 4 novel ecosystems, 350
challenges of, 298–301
jointly-owned, 127 novocaine, 66
climate change and, 299–300
valuation of, 55–61 nutrient cycles, 33
degradation issues, 299
zoning and, 294–296 wetland, 68
funding, 300–301
natural selection O personnel, 300–301
allele frequency and, 173
oceanic diversity, 45–46 poaching and, 298
fundamental theorem of, 179
oceanic islands, invasive species on, trophy hunting and, 298
Nature Conservancy, The (TNC), 81,
135–136 passive restoration, 340–341
215, 216f, 284f, 367, 368f, 386f
oceans, acidification of, 122–124 pathogens, 37
nature reserves, 268t
oil spills, 394 payments for ecosystem services
NatureServe network of Natural
omnivores, trophic levels, 36–37 (PES), 328–329, 328f, 370
Heritage programs, 215, 216f
On the Origin of Species (Darwin), 166 peace parks, 227
near threatened (NT) category, 213
Oostvaardersplassen, 346–347, 347f Penan peoples, Borneo, 12
New Zealand
open-access resources, 54 peregrine falcons (Falco peregrinus),
extinctions, 136
option values, 60, 78–80, 82f, 87, 115f, 313
wetlands, 281
227–228 Periyar Tiger Reserve, India, 180
New Zealand kakapo, 238
orangutan (Pongo spp.), 242, 242f, Permian Period extinctions, 152f
newts, habitats for, 314
308 perverse subsidies, 57, 398
Index 473

pest species, movement along cor- population density in protected areas, 289
ridors, 286 disease transmission and, 144 preservationist ethic, 8
pesticides social stress and, 145 prey species, predators and, 35
control of invasive species with, population establishment programs, primary consumers, 34f
140–141 238–239 trophic levels, 36
pollution by, 113–114 population sizes primary producers, 34f, 36
pharmaceuticals, natural, 66–67 death rates and, 185 primates, trade in, 129t
phenotypes, definition of, 32 declines, 124t private goods, 60
Philippines, destruction of reefs, determination of effectiveness, private lands
104–105 179–184 listed species on, 308
Phoenix Island Protected Area, fluctuations and bottlenecks, management of, 312–313
Kiribati, 272 181–183 procaine, 66
physical geography, 6 genetic diversity and, 171–176 productive use values, 61, 63–67, 64f
physiology, population biology and, genetic variation and, 174 Project Budburst, 394
194–195 loss of alleles and, 175 Project Nestwatch, 200
phytoplankton, 45f metapopulations and, 210f protected areas, 264–303
Pielou evenness index, 30 monitoring populations, 199–200 adjacent lands, 281
Pinchot, Gifford, 8 reduced, 108–109 characteristics, 280–283
Pitcairn Islands, 272 vulnerability to extinction and, classification of, 266–277
Pitcher’s thistle (Cirsium pitcheri), 167 conservation outside, 304–335
245f population studies, 195–203 definition of, 265
Plant Conservation (IUCN), 220 African parks, 209f designations, 268t
plant ecologists, field work, 26f fieldwork, 197–198 designing networks of, 277–283
plants long-term monitoring, 210–212 effectiveness of, 272
establishment programs, 243–244 published literature, 195–196 establishment of, 266–277
microsite conditions, 244 unpublished literature, 196–197 gap analyses, 274–277
predation on, 35 population viability analysis (PVA), local peoples and, 399–400
PlantSearch database, 258 203–208, 204f management of, 288–298
poaching, 298 populations monitoring, 292–294
polar ice caps, 123, 124t description of, 31 networks of, 283–286
polar regions, warming of, 120, 121f establishment of, 236–246 people in management of, 294
policy makers, collaboration among, monitoring of, 196f resources outside of, 305
17 reinforcement of, 236–246 sizes, 280–283, 305
pollination, bees in, 73f Possession Island, Crozet Archi- species migration between, 305
pollution pelago, 198 transfrontier parks, 396
air, 117–118 power line corridors, 311 types of, 267–268
bioremediation of, 348 prairie ecosystems, 70f visitors to, 289f
Convention on Long-range Trans- prairies world percentages, 269f
boundary Air Pollution, 381 plant communities, 344 protected landscapes, 268t
Convention on the Prevention of remnants, 311 protected seascapes, 268t
Marine Pollution by Dumping restoration projects, 355–357 provisioning services, 60–61
of Wastes and Other Matter, 383 Preble’s meadow jumping mouse Przewalski’s horse (Equus caballus
coral reefs and, 104–105 (Zapus hudsonius), 223 przewalski), 247–248, 248f
endangered species and, 97 precautionary principles, 57 pseudoephedrine, 66t
environmental degradation and, precipitation Public Broadcasting Services, 404
112–118 rain, evaporation of, 33 public goods, 68
environmental threats, 374 precipitation, acid rain, 117–118 Puno, Peru, 400f
international scope of, 228 predation, species interactions, 35, Pyrenean ibex (Capra pyrenaica pyre-
marine, 382 134 naica, bucardo), 261
pesticides, 113–114 Predator Free New Zealand pro- Q
water, 114–117 gram, 136
quagga mussels (Dreissena rostrifor-
polymorphic genes, 32 predator release hypothesis, 139
mis bugensis), 137, 138f
ponds, artificial, 292, 345 predators
Quaternary Period extinctions, 152f
population biology, applied, 194–212 insect, 344
population bottlenecks, 182 keystone species, 38–39
474 Index

Queensland Rain Forest, Australia, representative sites, 218 ScienceDirect, population studies,
380 reproduction rates, 169, 181 196
quinine, 66t reptiles scientific values, 77
QUINTESSENCE Consortium 2016, maintained in zoos, 250t scientists, collaboration among, 17
401 trade in, 129t scopolamine, 66t
R research, reserves and, 399 sea grasses, 311
reserpine, 66t sea horses (Hippocampus spp.), 129
rabbits
resilience, 41 sea levels, rising, 124t
Australian invasion by, 139
resiliency, 277 sea turtles, conservation of, 256
population growth, 188
resistance, 41 seabirds, decline of, 171
radiotelemetry, use of, 198
resource conservation ethic, 9 seasonal changes, 124t
rain, evaporation of, 33. see also acid
resource extraction, 283 secondary consumers, 34f, 36
rain
resources, competition for, 134 secondary invasions, 344
rain forests, species losses, 165
restoration ecology, 336–361 sedimentation, coral reefs and,
rainbow trout (Oncorhynchus mykiss),
approaches to, 340f 104–105
140
definition of, 337–338 seed banks, 258–261, 259f
Ramsar Convention on Wetlands,
degraded ecosystems, 341–342 seed dispersal, 108, 344
375f, 379, 379t
future of, 359 seed production, reduced, 108
rapid biodiversity assessments
moving targets of, 350 sei whale, 131t
(RAPs), 218–219
Restoration Ecology, journal of, 359 sennoside, 66t
rattlesnakes, 67
restoration remedies, 371 Sequoia and Kings Canyon National
reality, 278
rewilding, 346 Park, 220f
recombination, genetic variation
Rheobatrachus silus (gastric-breeding Serengeti National Park, Tanzania,
and, 32
frog), 261 380
reconciliation ecology, 306, 345
rhinoceroses, captive breeding, 253 sex ratios, unequal, 180
recovery criteria (RC), 206
Rio Declaration on Environment and shade coffee, 316f
recovery programs, 245
Development, 376 Shannon diversity index, 30
recreational activities, 97
Rio Earth Summit, 384f Shannon-Wiener index, 30
red-cockaded woodpecker (Picoides
Rio Summit, 376 Shepard, Paul, 86–87
borealis), 310
roads shifting cultivation, 100
Red Data Books, 215
endangered species and, 97 shrimp, pollution threats to, 115
Red List Index, 215
habitat fragmentation and, 108 Siberian ibex (Capra sibirica), 407
Red Lists (IUCN), 215
highway overpass, 285f Sierra Club, 81, 386f
Red Lists (IUCN) criteria, 213, 213t
Rocky Mountain spotted fever, 315 Silent Spring (Carson), 9
Red River wetlands, 310
rose periwinkle (Catharanthus roseus), Silurian Period extinctions, 152f
Red Sea Marine Peace Park, 382
66 Simpson index, 30
red-tailed hawk, 314f
Rosenzweig, Michael, 306 Singapore, 165, 282
red wolf (Canis rufus), 26
Roundup Ready soybeans, 142 Sinharaja Forest Reserve, Sri Lanka,
Reducing Emissions from Defores-
Royal Botanic Gardens, Kew, Eng- 380
tation and Forest Degradation
land, 256, 257f sink populations, 209
(REDD), 389
Royal Society for the Protection of sixth extinction episode, 153
redundancy, 277
Birds, 11, 367 skyscrapers, wildlife among, 313–314
reef corals, trade in, 129t
S slash-and-burn agriculture, 100
reference sites, 339
SLOSS (single large or several small)
Regional Seas Program, UN, 383 Sahel region, Africa, 106
debate, 280
regionally extinct (RE) category, 214 salamanders, hybrid, 27
small populations. see also popula-
regulating services, 61 salmon (Oncorhyncus spp.), 115, 170f
tion sizes
rehabilitation, degraded ecosystems, San Diego Zoo Institute for Conser-
genetic variability within, 174
341 vation Research, 254
Smith, Adam, 54
reinforcement programs, 236 Sandoz, 79
Smithsonian Conservation Biology
reintroduction programs, 236 Sarus crane (Grus antigone), 370f
Institute, 254
plant case studies, 244 satellite tags, use of, 198, 198f
Smithsonian Institution, 384
species selection, 239 scarlet gilia (Ipomopsis aggregata), 176
Sochi National Park, Western Cau-
replacement cost approaches, 63 scat, DNA analysis of, 201–202
casus, 274
representation, 277 scavenging behaviors, 36–37
Index 475

social behaviors, captive breeding Stephen’s kangaroo rat (Dipodomys timber, overharvesting, 69. see also
and, 243 stephensi), 226 logging
social justice issues, 364 stochasticity, definition of, 184 Tokyo, Japan, landscapes, 287t
social stress, disease and, 145 storks (Ciconia ciconia), 313 Tolypocladium inflatum, 79
Society for Conservation Biology, 12, subsistence farming Toronto, Canada, 94–95
403–404 shifting cultivation and, 100 tourism, biodiversity and, 16f
Society for Ecological Restoration, sulfur oxides, 117–118 towns, development of, 367
359 sulfuric acid, 117–118 toxic metals, 118
sociology, methods of, 6 Sumatra, deforestation and, 100, 102 toxic pollutants, breakdown of, 72
soft release approaches, 239–240 Sumatra, Indonesia, 158 toxiferine, 66t
soil, erosion of, 70–71 Sumatran orangutan (Pongo abelii), traditional people, conservation
soil organisms, 74 242f beliefs, 321–330
soil protection, water and, 70–72 sun coffee, 316f TRAFFIC network (WWF), 228
songbirds, migratory, 109 surveys. see also censuses transect lines, 201
Soulé, Michael, 12, 346 description of, 201–203 transfrontier parks, 396
source populations, 209 population monitoring with, 199 trapping, control of invasive species
South Africa sustainable development with, 140–141
Cape region, 46f challenges of, 362–391 Tree of Life Web Project, 396
ecological footprint of, 95f conservation biology and, 9 trees, importance of, 72
South Korea, demilitarized zone, 312 description of, 9 Triassic Period extinctions, 152f
southern right whale, 131t international approaches to, tribal people. see traditional people
soybeans, Roundup Ready, 142 374–383 trophic cascades, 39
Spanish imperial eagle (Aquila adal- local level, 365–372, 399–400 trophic levels
berti), 186 Svalbard Global Seed Vault, 258 biologic communities, 36
species swidden, 100 field ecosystem, 34f
definitions of, 25–28 symbiotic relationships, 35–36 trophic rewilding, 346
establishment rates, 163 Syria, protected areas, 268 trophy hunting, 298, 326
evolutional definition of, 26f tropical deciduous forests, 102
T
gap analysis of protection, 275f tropical forests, 45
tallgrass prairies, grazing in, 290f
identification of, 25 tropical rain forests
Tallis, Heather, 402
interdependence, 85 destruction of, 99–102
tamarisk (salt cedar, Tamarix spp.),
invasiveness of, 137–140 restoration projects, 357–359
134, 348, 349f
legal protection of, 222–231 species in, 99
Tamarisk Coalition, 349–350
morphological definition of, 25 species loss in, 159
tamarisk leaf beetle (Diorhabda spp.),
over geologic time, 152f surface area of, 99, 99f
348, 349f
relationships between, 73–74 tropical wetland ecosystems, 82f
Taoism, conservation and, 7
right to existence of, 84 trout stream habitats, 338f
tarantulas, 43
valuation of, 60 tubocurarine, 66t
Tasmanian tiger-wolf (Thylacinus
worldwide, 41–44, 42f Tukano Indians, Brazil, 322
cynocephalus), 154
species-area relationships, 162f, 164f turtle excluder devices (TEDs), 14
taxonomists, 25
species distribution, 47–48 Tuttle, Merlin, 405
tea plantations, 101f
global warming and, 124–125 Twitter, 405
terrestrial environments, 72
species diversity, 23–31, 24f
tertiary consumers, 36 U
species richness, 28, 47
Tertiary Period extinctions, 152f umbrella species, 217
Species Survival Commission
THC, 66t UNESCO Man and the Biosphere
(IUCN), 218, 253
Thermus aquaticus, 67 Program, 379t, 380–381
sperm whale (Physeter macrocepha-
Thomson Reuters, population stud- United Arab Emirates, ecological
lus), 131t
ies, 196 footprint of, 95f
spiders, species, 43
Thoreau, Henry David, 8, 158 United Kingdom
Spine of the Continent Initiative, 285
threatened species habitat modification in, 97
spiritual values, natural areas and, 8
animals, 153t protected areas, 270, 282f
St. Helena ebony (Trochetiopsis
plants, 153t United Nations Conference on Cli-
ebenus), 154, 154f
Three Gorges Dam, 103 mate Change, 13f
stakeholders, in ecosystem manage-
ment, 319, 319f, 344
476 Index

United Nations Conference on Waterton-Glacier International Peace Wildlife Conservation Society, 370f,
Environment and Development Park, 382 385
(UNCED), 376 Waterton Lakes National Park, wildlife management, benefits of,
United Nations Development Pro- Canada, 382 332
gramme, 385 weather, extreme events, 122 Wildlife Trade Program, 228
United Nations Educational, Scien- Web of Science, population studies, Wilson, E. O., 3, 359
tific and Cultural Organization 196 wind turbines, 57
(UNESCO), 294 Western lowland gorillas (Gorilla withering shell syndrome, 160f
biosphere reserves, 296–298 gorilla gorilla), 204–205 Wollerni pine (Wollemia nobilis), 176
United Nations Environment Pro- wetland ecosystems Wolong National Nature Reserve,
gramme (UNEP), 226, 383 coastal, 103–104 253f
United Nations Framework Conven- Convention on Wetlands, 379t wolves (Canis lupus) reintroduction,
tion on Climate Change (UN- economic values, wetlands, 82f 346
FCCC), 375f, 376 functions of, 72 woodlands, species diversity and,
United States functions vital to humans, 71f 288
control of invasive species, indirect use values, 68 World Bank, 383–386
140–141 interconnected, 292 World Commission on Environment
ecological footprint of, 95f Louisiana, 351 and Development, 374
funding wildlife conservation, 59 maintenance of, 292 World Conservation Monitoring
Kemp’s Ridley sea turtle recovery New Zealand, 281 Centre (WCMC), 229
efforts, 14 Ramsar Convention on, 375f World Heritage Convention (WHC),
resource consumption in, 4, 94 Red River, 310 375f, 379, 379t, 380
urban areas restoration of, 69, 351–353 World Heritage Sites, 300, 380, 381f
conservation in, 313–318 threats to, 103 World Network of Biosphere Re-
process alteration in, 307f whale watching, 81f serves, 293–294
restoration projects, 344–346 whales World Summit on Sustainable Devel-
Uruguay, protected areas, 268 extinctions, 159 opment, 375, 378, 378f
use values, 60 hunting of, 130 World Wildlife Fund (WWF), 81, 228,
utility, 217 protection of, 404 385, 386f
worldwide populations of, 131t World Zoo Conservation Strategy,
V
white abalone (Haliotis sorenseni), 252
Valdivan Coastal Reserve, Chile, 284f
160f
venomous animals, drugs from, 67 Y
white-footed mouse (Peromyscus
Vera, Frans, 346–347 Yangtze river dolphin (Lipotes vexil-
leucopus), 110, 145f
viability analysis, metapopulations, lifer), 155–156
white nose syndrome, 143–144, 144f
245f Yellowstone National Park, 67f
White Sands Missile Range, New
vincristine, 66t gray wolves in, 236, 237f
Mexico, 309–310
vulnerable species (VU) category, wolf reintroduction projects, 346
whooping crane (Grus americana),
158, 166–172, 213 Yosemite Grant Act, 265
59, 223
YouTube, 405
W Wikipedia, population studies, 196
Yunnan Province, China, 300
Walden (Thoreau), 8 Wild Kratts, 404
Wall-E, 404 wilderness approach, 218–219 Z
warfarin, 66t Wilderness Areas, 372 zebra mussel (Dreissen polymorpha),
water hyacinth (Eichornia crassipes), description of, 268t 137, 137f
137, 139f local residents and, 321–330 Zinave National Park, 382f
waters process retention in, 307f Zona Protectora Las Tablas, 285
cycles, 33 wildflowers, 158 zoning, 294–296
endangered species projects, 97 shade-tolerant, 111 Zoological Information Management
pollution of, 114–117 wildlife, trade in, 128–130, 129t System (ZIMS), 253
soil protection and, 70–72 Wildlife and Countryside Act, 222 zoos
unprotected, 311–312 “wildlife certification program,” goals of, 249–255
wetlands and purification of, 68 backyard, 312 terrestrial vertebrates in, 250t, 251f
watershed areas, 367
About the Authors
Richard B. Primack is a Professor of Biology at Boston University. He re-
ceived his B.A. at Harvard University in 1972 and his Ph.D. at Duke Univer-
sity in 1976, and was a postdoctoral fellow at the University of Canterbury
and Harvard University. He served as a visiting professor at the University
of Hong Kong, Tokyo University, and the Northeast Forestry University in
China, and has received Harvard’s Bullard and Putnam Fellowships, a Gug-
genheim Fellowship, and Germany’s Humboldt Fellowship. Dr. Primack
was President of the Association for Tropical Biology and Conservation, and
is currently Editor-in-Chief of the journal Biological Conservation. Thirty-four
foreign-language editions of his conservation textbooks have been produced
with local coauthors. He is an author of Tropical Rain Forests: An Ecological
and Biogeographical Comparison (with Richard Corlett). Dr. Primack’s research
interests include climate change, the loss of species, tropical ecology, and
conservation education. He has recently completed a popular book about
the impacts of climate change, titled Walden Warming: Climate Change Comes
to Thoreau’s Woods.
Anna A. Sher is a Professor of Biology at the University of Denver, where
she has taught Conservation Biology since 2003. She held a joint position
as the Director of Research and Conservation at Denver Botanic Gardens
from 2003–2010. Dr. Sher has published books and articles for academic,
trade, and popular audiences on various topics within conservation biology,
including restoration ecology, rare plant conservation, and climate change.
She is one of the foremost experts on the ecology of invasive Tamarix trees
and was the lead editor of the book Tamarix: A Case Study of Ecological Change
in the American West (Oxford University Press, 2013). Dr. Sher received her
Ph.D. in Biology at the University of New Mexico in 1998, and was a post-
doctoral fellow at the University of California, Davis and a Fullbright Schol-
ar in Israel. She has taught as adjunct faculty at both a small liberal arts
college, Earlham College, and a large state school, the University of New
Mexico. Dr. Sher also led scientific study-abroad programs in East Africa,
and is a contributing science writer for the Huffington Post blog.

About the Book


Editor: Rachel Meyers
Project Editor: Martha Lorantos
Production Manager: Christopher Small
Book Design: Joanne Delphia
Cover Design and Book Layout: Ann Chiara
Photo Researcher: David McIntyre
Book and Cover Manufacture: RR Donnelley

You might also like