Conservation Biology in Sub-Saharan Africa - Richard B. Primack - E-Book

Conservation Biology in Sub-Saharan Africa E-Book

Richard B. Primack

0,0
6,99 €

oder
-100%
Sammeln Sie Punkte in unserem Gutscheinprogramm und kaufen Sie E-Books und Hörbücher mit bis zu 100% Rabatt.
Mehr erfahren.
Beschreibung


Easy to read, this lucid and accessible textbook includes fifteen chapters that cover a full range of conservation topics, including threats to biodiversity, environmental laws, and protected areas management, as well as related topics such as sustainability, poverty, and human-wildlife conflict. This rich resource also includes a background discussion of what conservation biology is, a wide range of theoretical approaches to the subject, and concrete examples of conservation practice in specific African contexts. Strategies are outlined to protect biodiversity whilst promoting economic development in the region.

Boxes covering specific themes written by scientists who live and work throughout the region are included in each chapter, together with recommended readings and suggested discussion topics. Each chapter also includes an extensive bibliography.
Conservation Biology in Sub-Saharan Africa provides the most up-to-date study in the field. It is an essential resource, available on-line without charge, for undergraduate and graduate students, as well as a handy guide for professionals working to stop the rapid loss of biodiversity in Sub-Saharan Africa and elsewhere.
 

Das E-Book können Sie in Legimi-Apps oder einer beliebigen App lesen, die das folgende Format unterstützen:

EPUB
Bewertungen
0,0
0
0
0
0
0
Mehr Informationen
Mehr Informationen
Legimi prüft nicht, ob Rezensionen von Nutzern stammen, die den betreffenden Titel tatsächlich gekauft oder gelesen/gehört haben. Wir entfernen aber gefälschte Rezensionen.



Conservation Biology in Sub-Saharan Africa

John W. Wilson and Richard B. Primack

https://www.openbookpublishers.com

© 2019 John W. Wilson and Richard B. Primack

This work is licensed under a Creative Commons Attribution 4.0 International license (CC BY 4.0). This license allows you to share, copy, distribute and transmit the text; to adapt the text and to make commercial use of the text providing attribution is made to the authors (but not in any way that suggests that they endorse you or your use of the work). Attribution should include the following information:

John W. Wilson and Richard B. Primack, Conservation Biology in Sub-Saharan Africa. Cambridge, UK: Open Book Publishers, 2019, https://doi.org/10.11647/OBP.0177

In order to access detailed and updated information on the license, please visit, https://doi.org/10.11647/OBP.0177#copyright

Further details about CC BY licenses are available at, https://creativecommons.org/licenses/by/4.0/

All external links were active at the time of publication unless otherwise stated and have been archived via the Internet Archive Wayback Machine at https://archive.org/web

Updated digital material and resources associated with this volume are available at https://doi.org/10.11647/OBP.0177#resources

Licencing for images obtained online, many of which underwent minor stylistic adjustments, were correct at the time of publication.

Every effort has been made to identify and contact copyright holders and any omission or error will be corrected if notification is made to the publisher.

The Lounsbery Foundation has generously contributed to the publication of this volume.

ISBN Paperback: 978-1-78374-750-4

ISBN Hardback: 978-1-78374-751-1

ISBN Digital (PDF): 978-1-78374-752-8

ISBN Digital ebook (epub): 978-1-78374-753-5

ISBN Digital ebook (mobi): 978-1-78374-754-2

ISBN XML: 978-1-78374-755-9

DOI: 10.11647/ OBP.0177

Cover photo and design by Anna Gatti.

This book is dedicated to our spouses Lesley Starke and Margaret Primack, who supported our efforts to write this book and our decision to provide it for free to students and researchers across Africa

Contents

List of Boxes xix

Foreword xxv

Preface xxvii

Scope xxvii

Taxonomy and the IUCN Red List categories xxviii

Organisation of the book xxix

Bibliography xxx

Acknowledgements xxxi

List of Acronyms xxxiii

1. What is Conservation Biology?

1.1. Conservation Biology is Still Evolving 3

1.2 The Role of Conservation Biologists 7

1.3 The Value of Scientific Methods 9

1.4 Environmental Ethics 12

1.4.1 Conservation biology’s ethical principles 16

1.5 Summary 18

1.6 Topics for Discussion 19

1.7 Suggested Readings 19

Bibliography 20

2. Introduction to Sub-Saharan Africa

2.1 Sub-Saharan Africa’s Natural Environment 24

2.2 History of Conservation in Sub-Saharan Africa 29

2.2.1 The 1800s and launching of formal conservation efforts 33

2.2.2 Conservation efforts after colonialism 36

2.3 Conservation in Sub-Saharan Africa Today 37

2.4 Ongoing Conservation Challenges 44

2.4.1 Persistent poverty 44

2.4.2 Obstructive mindsets 46

2.4.3 Weak governance/institutional structures 47

2.4.4 Skills shortages 51

2.4.5 Competing interests 52

2.5 Conclusion 53

2.6 Summary 53

2.7 Topics for Discussion 54

2.8 Suggested Readings 54

Bibliography 55

3. What is Biodiversity?

3.1 Species Diversity 62

3.1.1 What is a species? 64

3.2 Genetic Diversity 65

3.3 Ecosystem Diversity 69

3.4 Patterns of Biodiversity 71

3.4.1 Challenging species identifications 72

3.4.2 Implications of challenging species identifications 73

3.4.3 Measuring species diversity 79

3.4.4 How many species exist? 81

3.4.5 Where are most species found? 82

3.5 Summary 85

3.6 Topics for Discussion 86

3.7 Suggested Readings 86

Bibliography 87

4. Why Should We Protect Biodiversity?

4.1 Material Contributions 93

4.2 Regulating Services 96

4.2.1 Maintaining ecosystem stability 96

4.2.2 Maintaining ecosystem productivity 99

4.2.3 Climate regulation 99

4.2.4 Conserving soil and water quality 100

4.2.5 Pollination and seed dispersal 101

4.2.6 Hazard detection and mitigation 105

4.2.7 Pest and disease control 106

4.3 Nonmaterial Contributions 112

4.3.1 Inspiration and learning support 112

4.3.2 Supporting psychological and physical experiences 114

4.3.3 Supporting individual and group identities 116

4.4 The Long-Term View: Option Values 117

4.5 Environmental Economics 118

4.5.1 Placing a price on the natural world 119

4.5.2 Environmental economics’ biggest contributions 120

4.5.3 Environmental economics’ biggest challenges 121

Accounting for negative externalities 121

Determining ownership 122

A more inclusive approach 123

4.6 Summary 124

4.7 Topics for Discussion 124

4.8 Suggested Readings 125

Bibliography 126

5. The Scramble for Space

5.1 What is Habitat Loss? 134

5.1.1 What is habitat fragmentation? 138

5.1.2 What are edge effects? 140

5.2 Drivers of Habitat Loss and Fragmentation 142

5.3 Habitat Loss’ Impact on Africa’s Ecosystems 145

5.3.1 Tropical forests 145

5.3.2 Rivers and deltas 148

5.3.3 Wetlands 149

5.3.4 Seasonal drylands 153

5.4 Population Growth and Consumption? 156

5.5 Concluding Remarks 159

5.6 Summary 160

5.7 Topics for Discussion 160

5.8 Suggested Readings 161

Bibliography 161

6. Our Warming World

6.1 Drivers of Climate Change 168

6.2 Predicting Earth’s Future Climate 173

6.3 The Impact of Climate Change 175

6.3.1 Climate change’s impact on people 176

6.3.2 Climate change’s impact on terrestrial ecosystems 178

Climate change on mountains 178

Climate change in the lowlands 178

Climate change and dispersal limitations 182

Climate change and biological interactions 182

Climate change and reptiles 184

6.3.3 Climate change’s impact on freshwater ecosystems 184

Warmer rivers and streams 184

Changing flow regimes 185

6.3.4 Climate change’s impact on marine ecosystems 185

Ocean acidification 186

Sea level rise 187

Coral bleaching 187

Ocean deoxygenation 188

6.3.5 Climate change interacts with habitat loss 188

6.4 Beneficiaries of Climate Change 189

6.5 The Overall Impact of Climate Change 192

6.6 Summary 193

6.7 Topics for Discussion 194

6.8 Suggested Readings 194

Bibliography 195

7. Pollution, Overharvesting, Invasive Species, and Disease

7.1 Pollution in Its Many Forms 204

7.1.1 Water pollution 207

7.1.2 Air pollution 212

7.1.3 Soil pollution 213

7.1.4 Light pollution 214

7.1.5 Noise pollution 216

7.1.6 Thermal pollution 216

7.2 Overharvesting 218

7.2.1 The Bushmeat Crisis 218

7.2.2 Overfishing 220

7.2.3 The impact of traditional medicine 223

7.2.4 The impact of live animal trade 224

7.2.5 Overharvesting of plant products 226

7.2.6 Challenges in managing overharvesting 226

7.3 Persecution 230

7.4 Invasive Species 231

7.4.1 Spread of invasive species 231

7.4.2 Impact of invasive species 235

7.4.3 Genetically modified organisms 237

7.6 Parasites and Diseases 238

7.7 Summary 242

7.8 Topics for Discussion 243

7.9 Suggested Readings 244

Bibliography 244

8. Extinction Is Forever

8.1 What is Extinction? 259

8.2 Rates of Extinction 259

8.3 When is a Species Extinct? 260

8.4 History of Extinctions in Sub-Saharan Africa 263

8.5 Which Species are at Risk of Extinction? 271

8.5.1 Course-filter assessments 274

8.6 Characteristics of Threatened Species 274

8.7 Problems of Small Populations 277

8.7.1 Loss of genetic diversity 277

Genetic drift 277

Inbreeding depression 278

Outbreeding depression 280

Population bottlenecks 281

8.7.2 Demographic stochasticity 281

8.7.3 Environmental stochasticity and catastrophes 282

8.7.4 The extinction vortex 284

8.7.5 Is there any hope for small populations? 285

8.8 Is De-extinction a Solution? 285

8.9 Summary 289

8.10 Topics for Discussion 289

8.11 Suggested Readings 290

Bibliography 291

9. Applied Population Biology

9.1 Monitoring Population Size 299

9.1.1 Biodiversity inventories 299

9.1.2 Population censuses 302

9.1.3 Demographic studies 306

9.1.4 Recent progress in collecting survey data 308

9.2 Estimating Extinction Risk 309

9.2.1 A word of warning 310

9.2.2 Probability of extinction 310

9.2.3 Minimum viable population 311

9.2.4 Effective population size 313

9.2.5 Maximum sustainable yield 314

9.2.6 Sensitivity analysis 317

9.3 Challenges to PVA Implementation 318

9.3.1 Lack of adequate data 318

9.3.2 Data reliability 319

9.3.3 Model reliability 319

9.4 Summary 320

9.5 Topics for Discussion 320

9.6 Suggested Readings 321

Bibliography 322

10. Conserving Ecosystems

10.1 Ecosystem Monitoring 328

10.1.1 Monitoring ecosystems with geospatial analysis 332

10.2 Maintaining Complex and Adaptive Ecosystems 337

10.2.1 Maintaining critical ecosystem processes 338

The water cycle 338

The nutrient cycle 339

The energy cycle 340

Community dynamics 341

Fire Dynamics 341

10.2.2 Minimising external threats 343

Controlling invasive species 344

10.2.3 Adaptive management 349

10.2.4 Being minimally intrusive 353

10.3 Restoring Damaged Ecosystems 354

10.3.1 Ecological restoration approaches 355

10.3.2 Major restoration targets 357

10.3.3 The future of ecological restoration 362

10.4 Combatting Climate Change Through Ecosystem Conservation 363

10.5 Summary 364

10.6 Topics for Discussion 365

10.7 Suggested Readings 366

Bibliography 366

11. Preventing Extinctions

11.1 Studying Species and Populations 376

11.1.1 Obtaining natural history data 380

11.2 Saving Species Through Translocations 384

11.2.1 Important considerations for translocations 384

Determining need and feasibility 385

Support from local stakeholders 385

Identifying suitable habitat 386

Considering genetics and behaviour 388

How many individuals to release 388

Preparing individuals for release 391

Post-release monitoring 394

Helping other translocation projects 394

11.3 Managing and Facilitating Movement Dynamics 395

11.3.1 Connectivity in terrestrial ecosystems 396

11.3.2 Connectivity in freshwater ecosystems 402

11.3.3 Connectivity in marine ecosystems 402

11.3.4 Mimicking connectivity 403

11.3.5 Management considerations in connectivity conservation 403

11.4 Managing Species Sensitive to Climate Change 404

11.5 Ex Situ Conservation Strategies 406

11.5.1 Types of ex situ facilities 409

11.5.2 Challenges facing ex situ facilities 411

11.6 Thoughts on Neglected Taxa 413

11.7 Summary 417

11.8 Topics for Discussion 418

11.9 Suggested Readings 418

Bibliography 419

12. Biodiversity and the Law

12.1 Identifying Legislative Priorities 428

12.2 Environmental Laws and Policies 430

12.2.1 International agreements 430

12.2.2 National and local laws 435

12.3 Environmental Law Enforcement 439

12.3.1 New technologies in environmental law enforcement 440

12.4 The Limits of Environmental Laws and Regulations 448

12.4.1 Lack of capacity 449

12.4.2 Conflicting government priorities 449

12.4.3 Informal economies, traditional activities, and the law 450

12.4.4 Trade embargoes and sanctions 451

12.5 Conclusion 452

12.6 Summary 455

12.7 Topics for Discussion 455

12.8 Suggested Readings 456

Bibliography 457

13. The Importance of Protected Areas

13.1 Establishing Protected Areas 462

13.1.1 Government protected areas 463

13.1.2 Community conserved areas 464

13.1.3 Privately protected areas 464

13.1.4 Co-managed protected areas 465

13.1.5 Field stations and marine laboratories 466

13.2 Classification of Protected Areas 466

13.3 Prioritisation: What Should be Protected? 469

13.3.1 Species approach 471

13.3.2 Ecosystem approach 471

13.3.3 Wilderness approach 472

13.3.4 Hotspot approach 472

13.3.5 Gap analysis approach 474

13.3.6 Optimisation approach 478

13.4 How Much Land Should We Protect? 478

13.4.1 A neglected system: marine protected areas 480

13.5 Designing Protected Areas 480

13.5.1 What size should a protected area be? 483

13.5.2 Zoning as a solution to conflicting demands 486

13.5.3 Connectivity among protected areas 492

13.5.4 What about small isolated reserves? 493

13.6 Managing Protected Areas 494

13.6.1 The importance of monitoring 495

13.6.2 The importance of working with local people 496

13.6.3 The importance of accommodating visitors 498

13.6.4 The IUCN Green List of Protected Areas 499

13.7 Challenges for Protected Areas 500

13.7.1 Funding limitations 501

13.7.2 Planning for climate change 502

13.7.3 Facing degazettement 502

13.8 Summary 503

13.9 Topics for Discussion 504

13.10 Suggested Readings 504

Bibliography 505

14. Conservation on Unprotected Lands

14.1 Human-Dominated Landscapes 516

14.1.1 The impact of agriculture 525

14.1.2 The impact of logging, mining, and other extractive industries 530

14.2 Smart Development Outside Conservation Areas 532

14.3 Linking Conservation to Socio-Economic Development 534

14.4 Confronting Human-Wildlife Conflict 539

14.4.1 Dealing with predators 540

14.4.2 Dealing with crop raiders 541

14.4.3 Concluding thoughts on human-wildlife conflict 542

14.5 Summary 542

14.6 Topics for Discussion 543

14.7 Suggested Readings 544

Bibliography 545

15. An Agenda for the Future

15.1 Achieving Sustainable Development 556

15.2 Dealing with Technological Advances 558

15.3 Funding Conservation Activities 562

15.3.1 How effective is conservation funding? 564

15.4 Building Lasting Partnerships 568

15.4.1 Partnerships with local people 569

15.4.2 Partnerships among conservation professionals 573

15.5 Environmental Education and Leadership 574

15.6 Summary 579

15.7 Topics for Discussion 580

15.8 Suggested Readings 580

Bibliography 581

Appendix A 589

Selected Sources of Information

Appendix B 593

Selected Environmental Organisations

Appendix C 607

Obtaining Conservation Funding

Appendix D 613

Environmental Calendar

Glossary 615

Index 645

List of Boxes

Box 1.1 Conservation Through Public Health: A Case Study 5

Gladys Kalema-Zikusoka

Box 1.2 The Okapi Wildlife Reserve: Protecting Nature and Providing for People 14

Rosmarie Ruf & Marcel Enckoto

Box 1.3 Biodiversity: Can Humanity be Saved? 17

Nkengifor Nkeshia Valery

Box 2.1 Sacred Spaces: A Tradition of Forest Conservation in Benin 30

Emile N. Houngbo

Box 2.2 Why Go Transfrontier? (And Why Not?) 38

Tamar Ron

Box 2.3 Privately Owned Lands for African Conservation 40

Graeme Cumming

Box 2.4 Malawi: No Longer a Weak Link in the Elephant Ivory Trafficking Chain? 48

Jonathan Vaughan

Box 3.1 Finding a Needle in a Haystack: Monitoring Species Using eDNA 66

Tammy Robinson and Clova Mabin

Box 3.2 Golden Mole Conservation Requires a Sound Taxonomy 74

Sarita Mareeand Samantha Mynhardt

Box 3.3 Does Tardy Recognition of a Species Hamper its Conservation? 78

Colleen T. Downs

Box 4.1 Research on Hunting Underpins Conservation in Central Africa 94

Katharine Abernethy and Lauren M. Coad

Box 4.2 Are Wild Pollinators Important in African Agriculture? 103

Abraham J. Miller-Rushing

Box 4.3 Biological Control Saves the Cassava Crop 107

Meg Boeni and Richard Primack

Box 4.4 Conservation Lessons from the Asian and African Vulture Crises 110

Ara Monadjem

Box 5.1 The Importance of Liberia’s Forest Network to the Survival of the Pygmy Hippopotamus 136

Mary Molokwu-Odozi and Kathryn Phillips

Box 5.2 The Conservation and Exploitation of East African Plants 146

John R. S. Tabuti

Box 5.3 Migratory Birds of Africa: The Largest of the Last Great Migrations? 150

Abraham J. Miller-Rushing and John W. Wilson

Box 5.4 Saving Critically Endangered Ground Nesting Birds from Habitat Loss 153

Bruktawit Abdu Mahamued

Box 6.1 Does Oil Palm Agriculture Threaten Biological Diversity in Equatorial Africa? 170

Abraham J. Miller-Rushing

Box 6.2 Desert Birds and Climate Change 179

Susan Cunningham and Andrew McKechnie

Box 6.3 Habitat Alteration, Climate Change, and Mosquito-Borne Diseases 190

Kevin Njabo

Box 7.1 Solving Seabird Bycatch Problems: From Theory to Practice 221

Ross Wanless

Box 7.2 Conserving Elephants in the Anthropocene 227

David H.M. Cumming

Box 7.3 Aliens on Islands: Damage and Control 233

Peter Ryan

Box 7.4 Promoting African and Global Honeybee Health 238

Vincent Dietemann

Box 8.1 Pleistocene Extinctions: Climate Change, Hominin Predation, or Both? 263

David H.M. Cumming

Box 8.2 Swimming Dangerously Close to Extinction: Population Crash in Lesotho’s Endemic Maloti Minnow 269

Jeremy Shelton

Box 8.3 Fenced Reserves Conserving Cheetahs and African Wild Dogs in South Africa 286

Kelly Marnewick

Box 9.1 The Role of Biodiversity Inventories in the Management of Gorongosa National Park 300

Marc Stalmans and Piotr Naskrecki

Box 9.2 Sea Turtle Conservation along Africa’s Atlantic Coast 303

Angela Formia

Box 9.3 Sustainably Harvesting Fruit Bats Through Better Understanding of Life Histories 315

David T. S. Hayman

Box 10.1 Using Insects to Monitor Environmental Health 331

Rosina Kyerematen

Box 10.2 Remote Sensing and Spatial Analysis for African Conservation 335

Barend F. N. Erasmus

Box 10.3 Environmental Governance in the Serengeti Ecosystem 351

Alex Wilbard Kisingo

Box 10.4 Sustainable Forest Restoration Using Natural Vegetation 357

Samuel Kiboi

Box 11.1 The Overlooked Role of Behavioural Ecology in the Conservation of African Mammals 378

Adrian M Shrader

Box 11.2 Large Predator Reintroductions: A Balancing Act 389

Craig J. Tambling

Box 11.3 Transfrontier Conservation Areas: Managing Biodiversity Across International Boundaries 398

Simon M. Munthali

Box 11.4 Saving the Northern White Rhinoceros with Assisted Reproduction Technologies 414

Morné de la Rey

Box 12.1 Insect Biodiversity Helps Solve African Wildlife Crimes 442

Martin H. Villet

Box 12.2 Protecting Elephants in a Hostile Region 444

Lorna Labuschagne

Box 12.3 Thoughts on Poaching and Illegal Wildlife Trafficking in Sub-Saharan Africa 453

Tamar Ron

Box 13.1 Mpala Research Centre: A Living Laboratory for (More than Just) Scientists 467

Anchal Padukone and Dino J. Martins

Box 13.2 Identifying Key Sites for Conservation in the Albertine Rift 475

Andrew J. Plumptre

Box 13.3 Marine Protected Areas in East Africa and the Western Indian Ocean 481

Abraham J. Miller-Rushing

Box 13.4 Zoning: Something for Everyone in the Forests of Dzangha-Sangha 487

Richard Carroll

Box 14.1 Traditional People and Conservation: Turning the Page 517

Abraham J. Miller-Rushing and John W. Wilson

Box 14.2 Importance of Protected Areas in Cities: Insights from the City of Cape Town 523

Pippin M. L. Anderson

Box 14.3 Preserving Biodiversity Through Shaded Agroforestry 528

Hervé D. Bisseleua

Box 14.4 Confronting Human-Wildlife Conflict in Zimbabwe 537

Steven Matema

Box 15.1 Not Just for War: Drones in Conservation 559

Meg Boeni and Richard Primack

Box 15.2 Supporting Self-Organised Action for Conservation in Africa 565

Duan Biggs

Box 15.3 Tracking Species in Space and Time: Citizen Science in Africa 570

Phoebe Barnard

Box 15.4 The Contribution of Education Towards Conservation in Africa 576

Shiiwua Manu and Samuel Ivande

Foreword

Even for those of us who are one step removed from nature in our present-day lives, looking back on places we called home as children can reveal disturbing truths. This is certainly true for me. I grew up in a small mining town in Zimbabwe called Kadoma, and as soon as school was out, my parents would bundle me off to Mhondoro, a rural area to the east.

I loved it there and happily spent hours herding cattle across the quintessential untamed African savannah. I have many fond memories of swimming and catching fish in Mhondoro too. I recall how the surrounding forests were lush, teeming with wildlife; rivers were bountiful, full of fish and fowl in the pristine fresh water. City folk went home with their hands full of gifts from nature and agricultural fields.

Today, when I return to my village in Mhondoro, my heart breaks. The lush forests and wildlife are gone, replaced with barren fields and a whimpering stream where the river once ran. I now bring my own food and bottled water when I visit. And worst of all, the people of Mhondoro who I had always associated with nature’s abundance are today poor and disenfranchised and have few if any options for bettering their lives.

Tragically, the story of my village is shared by thousands of villages across Africa that are suffering the worst impacts of climate change, population growth and harmful development choices. Faced with the challenge of feeding their families and generating cash incomes, farmers, like those of Mhondoro, end up expanding their crops increasingly deeper into wild lands and forests. These encroachments not only bring their families into dangerous conflict with wildlife, they simultaneously endanger and destroy the forests and fertile soils that would otherwise support their agricultural bounty.

But this outcome is not inevitable.

Africa suffers the greatest burden of global heating and deteriorating nature. As such, there is recognition that a “new deal for nature” is needed if we are to avert the worst climate and nature crisis. A new deal that transforms the way we produce our food and choose what to consume, the way we develop infrastructure, including our cities, roads, housing and dams, produce our energy and the way we value nature in our economic systems. The search is on for solutions.

But Africa risks being left behind or having to acquiesce to solutions that are not fit nor ideal for the continent. If Africa is to meaningfully define solutions for a new deal for nature, we must support research capacity and skills building within its populations, including investments in faculty and research leaders, facilities and infrastructure, and expanding career opportunities for budding researchers to apply their findings in real world settings.

Fortunately, there is now widespread acknowledgement that African researchers are best placed to ask questions and find solutions to the challenges facing Africa. Hence, we are witnessing a slow move away from the notion that researchers from high income countries must be parachuted in to identify and address the continent’s problems.

This textbook, Conservation Biology in Sub-Saharan Africa, is a critical first step as it goes a long way towards focusing attention to the urgent need to define the nature of the problem and develop practical, context specific solutions to Africa’s many environmental challenges. It discusses how our lives are inextricably linked to a healthy environment, explores how inclusive conservation can provide greater benefits to all, and illustrates how grassroots action can ensure that nature’s many beneficial contributions will remain available for generations to come.

By being distributed for free, this textbook ensures that its readership can include those who stand to benefit the greatest, including African researchers and practitioners. Open access textbooks like it are critical to expanding access to African research and improving intra-African research collaboration and capacity.

If we are to “leave no one behind,” as agreed by global leaders in 2015 and encapsulated in Sustainable Development Goals, farmers like those from Mhondoro will also need access to the best available information, science and solutions. That is a deal that we owe them—and the many people who call Africa their home. This book makes an important contribution to the challenge. I hope that, thanks to the efforts of the experts featured in this textbook and others, one day when I return to Mhondoro, it will more closely resemble its prior self that I loved as a child.

Maxwell Gomera

Director: Biodiversity and Ecosystem Services Branch

UN Environment

Preface

We are excited to present the very first conservation biology textbook dedicated entirely to an African audience. The need for this work has never been more pressing than now. Africa has some of the fastest growing human populations on Earth. This growth, together with a much-needed push for development to ensure that all Africans can live healthy and prosperous lives, exerts unsustainable pressure on the region’s rich and unique biological treasures. Consequently, Africa is losing its natural heritage faster than ever before. It is sobering to consider that there is a very real risk that our children may never have the opportunity to see gorillas, rhinoceros, or elephants in the wild.

To address this alarming loss of Africa’s natural heritage, there is an urgent need to produce the next cohort of well-trained conservation leaders, able to confront conservation challenges head-on, and to secure a sustainable future for all. This effort starts early, by exposing children from a young age to the wonders of the natural world. But it is also important to ensure that those children who later choose a career in biodiversity conservation are well-prepared for the road ahead. To facilitate this capacity building, we have compiled this textbook, designed for use in conservation biology courses, and as a supplemental text for other courses in the natural sciences and environmental policy. While the main target audience for this book is early-career conservationists, we strived for a balance between theory, empirical data, and practical guidelines to also make the book a valuable resource for mid- and late-career professionals. To further remove obstacles to training, we made every effort to ensure that this work is accessible to as wide an audience as possible. For that reason, we are making this textbook available for free, under a Creative Commons (CC BY) license, to guarantee the rights for anyone to use and spread this work to whoever wishes to make a difference in the future of Africa’s biodiversity and its people.

Scope

This textbook focuses on the Afrotropics, one of Earth’s eight major terrestrial ecozones. This area includes continental Africa south of the Sahara Desert, continental islands (e.g. the Seychelles) that drifted away from Africa millions of years ago, and oceanic islands with a volcanic origin (e.g. the Comoros archipelago, São Tomé, and Príncipe) which share many biological characteristics with the Afrotropics. This area is also generally known as Sub-Saharan Africa, which we use throughout this book as a convenient and acceptable way to designate this ecologically (African parts of Afrotropics) and geographically (Africa south of the Sahara) distinct region.

Composite satellite image of Africa, with the Sahara Desert (sand-coloured) in the north and the Afrotropical ecoregion’s tropical ecosystems (in green) further south featuring prominently. The area between the sand-coloured and green regions is the Sahel, which marks the northern boundary of the Afrotropics. Photograph by NASA, https://commons.wikimedia.org/wiki/File:Africa_satellite_orthographic.jpg, CC0.

Deviating slightly from the typical scope of some other books focussed on the Afrotropics, this textbook does not cover south-west Arabia—these areas, together with North Africa, are covered in a sister textbook to this one, published in Arabic (Primack and El-Demerdash, 2003). Similarly, this book excludes Madagascar and the Mascarene Islands, which are covered in two sister textbooks published in French (Primack and Ratsirarson, 2005; Primacket al., 2012). The text does cover a few areas not usually considered part of the Afrotropics, but which share several affinities with the Afrotropical region. These additional areas include oceans within 200 nautical miles from Sub-Saharan Africa, and oceanic islands in the Atlantic that are usually treated as part of the Palearctic realm, namely Cabo Verde, St. Helena, Ascension, and the Tristan da Cunha archipelago.

Taxonomy and the IUCN Red List categories

The International Union for Conservation of Nature and Natural Resources (IUCN) maintains a comprehensive online database (https://www.iucnredlist.org) that summarises the threat status of many species on Earth. The classification system used to compile this database is discussed in detail in Section 8.5. We foreshadow this discussion by alerting readers that the threat status of each assessed species that is mentioned in the text is indicated with one of the seven acronyms mentioned below, right after its scientific name. To facilitate this indication, we generally follow the IUCN’s taxonomy in this textbook. We are fully aware that the lag time between taxonomic updates and IUCN assessment updates may create the appearance that this textbook’s taxonomy might sometimes be outdated. We made every effort to highlight important taxonomic discrepancies when relevant to the text. For common names, we tried to use the most-widely used terms across Africa.

EX

Extinct

EW

Extinct in the Wild

EN

Endangered

VU

Vulnerable

NT

Near Threatened

LC

Least Concern

DD

Data Deficient

Organisation of the book

This book contains 15 chapters and four appendices. While there is broad overlap in the topics covered in each chapter, the first three chapters are meant to be introductory, while chapter four provides an overview on the importance of biodiversity for our own wellbeing. Chapters five to seven outline the most important threats to biodiversity, while chapters eight to fifteen suggest overarching solutions to the current biodiversity crisis. We ensured that the main body text, which is nearly entirely comprised of examples from Sub-Saharan Africa, covers examples from a range of organisms living in terrestrial, freshwater, and oceanic environments. Within the main body text several words are written in bold-face—these represent the first mention and/or explanation of specialist terms listed in the glossary. Also included are over 50 Boxes, authored by conservation researchers and practitioners from governments, universities, and nongovernmental organisations in West, Central, East, and Southern Africa. These case studies cover interdisciplinary topics such as public health, sacred spaces, energy, agriculture, law, sustainable development, and leadership. They expose readers to the voices of conservationists in the region, provide compelling examples of on-the-ground work, and offer insights into real-life conservation issues readers may face in their careers.

In each chapter, following the main body of text, there is a brief summary of main take-home messages, a list of discussion questions, and a list of suggested readings. The discussion questions are formulated in such a way that there are no definitive right or wrong answers—rather, they are meant to stimulate discussion among readers so that they can develop their own conservation philosophies. We tried to restrict the suggested reading lists to works that are freely available either on publisher websites or online depositories accessible through Google Scholar (https://scholar.google.com). Note that the suggested readings are not meant to be absolute; in fact, they should be adapted to meet local contexts and syllabus requirements. Each chapter concludes with an extensive bibliography, which serves to provide a starting point for readers interested in specific conservation topics, and lecturers interested in adapting the suggested reading list. The textbook concludes with four appendices, meant to encourage readers to take the field’s activist spirit to heart.

While we have made every attempt to ensure that the content is current and comprehensive, we recognise that mistakes do creep in; the field of conservation biology is currently also rapidly evolving, including being more inclusive of previously marginalised communities. We thus plan on continue updating this text, and welcome comments and suggestions from readers who share our interest in protecting Sub-Saharan Africa’s natural heritage.

Bibliography

Primack, R.B., and J. Ratsirarson. 2005. Principe de Base de la Conservation de la Biodiversité (in French) (Antananarivo: University of Antananarivo).

Primack, R.B., F. Sarrazin, and J. Lecompte. 2012. Biologie de la Conservation (in French) (Paris: Dunod).

Primack R.B., and M. El-Demerdash. 2003. Essentials of Conservation Biology (in Arabic) (Cairo: Mars Publications).

Acknowledgements

We sincerely appreciate the contributions of everyone who helped ensure that this project was a success. In particular, we would like to thank Peter Ryan for his guidance and support throughout the project, Abe Miller-Rushing for writing several boxes and providing editorial input, and the Richard Lounsbery Foundation for funding support. Peter Ryan, David Cumming, and Adrian Shrader provided invaluable reviews to ensure the content is accurate and pitched at the right level for conservation students. We would also like to recognise the support from Nigel Bennett, Duan Biggs, Meg Boeni, Israel Borokini, and Mike Scantlebury at various stages of the project. Additional input on ex situ conservation was provided by Dalia Amor Conde, Johanna Staerk, and Simeon Smeele at Species360, while Peter Maniloff provided input on environmental economics. We thank all our photographers and box authors for their willingness to contribute to this work. Alessandra Tosi, Rupert Gatti, and the rest of the staff at Open Book Publishers helped to transform the manuscript into a finished product. Lastly, a special thanks to Lesley Starke and Margaret Primack for their encouragement, support, advice, and input throughout the project.

List of Acronyms

CAR

Central African Republic

CBD

Convention on Biodiversity Diversity

CBNRM

community-based natural resource management

CITES

Convention on International Trade in Endangered Species of Fauna and Flora

DRC

Democratic Republic of the Congo, formerly Zaire

EIA

environmental impact assessment

FFI

Fauna & Flora International

FSC

Forest Stewardship Council

GDP

gross domestic product

GEF

Global Environmental Facility

GIS

geographic information system

GPS

Global Positioning System

FZS

Frankfurt Zoological Society

HIV

human immunodeficiency virus

IBA

Important Bird and Biodiversity Area

IPBES

Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services

IUCN

International Union for Conservation of Nature and Natural Resources

IPCC

Intergovernmental Panel on Climate Change

MPA

marine protected area

MSC

Marine Stewardship Council

MVP

minimum viable population

NGO

non-governmental organisation

PADD

protected area downgrading, downsizing, and degazettement

PAAZA

Pan-African Association for Zoos and Aquaria

PVA

population viability analysis

RLE

Red List for Ecosystems

RSPO

Roundtable on Sustainable Palm Oil

SCB

Society for Conservation Biology

SDM

species distribution modelling

sp.

species (plural: spp.)

TEK

traditional ecological knowledge

TFCA

transfrontier conservation area

TNC

The Nature Conservancy

UK

United Kingdom

UN

United Nations

UNEP

United Nations Environment Programme

UNESCO

United Nations Educational Scientific and Cultural Organisation

USA

United States of America

USAID

United States Agency for International Development

WCMC

World Conservation Monitoring Centre, a centre within UNEP

WCS

Wildlife Conservation Society

WWF

World Wide Fund For Nature

ZSL

Zoological Society of London

1. What is Conservation Biology?

© 2019 J.W. Wilson and R.B. Primack, CC BY 4.0 https://doi.org/10.11647/OBP.0177.01

1.1. Conservation Biology is Still Evolving p. 3

1.2 The Role of Conservation Biologists p. 7

1.3 The Value of Scientific Methods p. 9

1.4 Environmental Ethics p. 12

1.5 Summary p. 18

1.6 Topics for Discussion p. 19

1.7 Suggested Readings p. 19

Bibliography p. 20

East Africa’s great migration is one of the most famous wildlife spectacles on Earth. Each year, tens of thousands of tourists from around the world flock to the region to see the 1.7 million common wildebeest (Connochaetes taurinus, LC) and hundreds of thousands of other plains mammals make their way from Tanzania’s Ngorongoro Conservation Area, through the Serengeti Plains, to Kenya’s Maasai Mara National Reserve. Photograph by Daniel Rosengren, https://commons.wikimedia.org/wiki/File:Wildebeest_Migration_in_Serengeti_National_Park,_Tanzania.jpg, CC BY 4.0.

Popular interest in protecting biological diversity—which describes the amazing range of species, genetic diversity within each species, and the multitude of Earth’s complex biological communities with their associated ecosystemprocesses—has intensified during the past few decades. During this time, scientists and the public have recognised that biological diversity (often shortened to biodiversity) is being lost at increasing rates. Across the world, human activities are destroying ecological communities that have developed over millions of years. Over the next several decades, thousands of species and millions of populations will likely go extinct.

The fundamental driver of all the biodiversity losses we are currently witnessing is a rapidly expanding human population coupled with increased consumptive needs. In 1850, after roughly 300,000 years of Homo sapiens on the planet, there were around 1 billion people on Earth. By 1987, not even 140 years later, the world’s human population had surpassed 5 billion. By 2017, there were 7.5 billion humans globally, of which over 1 billion lived in Sub-Saharan Africa (World Bank, 2019). With this many people, the human population grows by tens of millions of people each year, even with modest population growth (Figure 1.1). To make matters worse, Sub-Saharan Africa has the fastest population growth rate in the world, with a projected human population estimate of over 4 billion people by the year 2100—a number that is well beyond the ecological capacity of the region to support.

Figure 1.1 Sub-Saharan Africa’s human population crossed the 1 billion mark in 2015. At the current annual population growth rate of 2.7%, more than 28 million people will be added to the region in 2019. This number will escalate each subsequent year as increases are compounded. Sources: Biraben, 2003; World Bank, 2019, CC BY 4.0.

To survive and prosper, people use natural resources. They harvest and use oil, water, and wildlife products, and convert natural ecosystems for agriculture, cities, roads, and industrial activities. This consumption, which reduces natural habitat and the associated wildlife populations, is intensifying because of the demands of a rapidly increasing human population. Consumption of resources also increases as countries develop and industrialise: the average citizen of the USA uses five times more resources than the average global citizen, 11 times more than the average Chinese citizen, and 32 times more than the average Kenyan citizen (Worldwatch Institute, 2015). This growth in the number of humans, together with their ever-more-intensive use of natural resources, is the fundamental driver behind most current species extinctions.

For conservation biologists and other nature lovers, the widespread extinction of species and destruction of natural ecosystems are incredibly discouraging.

For conservation biologists and other nature lovers, the widespread extinctions of species and destruction of natural ecosystems are incredibly discouraging. Perhaps nowhere in the world is this issue as dramatic as in Africa with its rich and spectacular wildlife, but also its significant socio-economic challenges, such as a rapidly increasing human population, persistent poverty, weak governance structures, and many people’s near-obligate dependence on natural resources. Many Africans are also confused by the importance and need for conservation actions, pointing to the romanticised but inaccurate notion that humans have been living in relative harmony with nature since humans first made an appearance on Earth (see Box 8.1). But it is possible, and indeed necessary, to find ways to ensure the persistence of biodiversity. Actions taken, or not taken, during the next few decades will determine how many species and natural areas will continue to survive. Someday, people will likely look back and say that this time—the first half of the 21st century—was an important and exciting time when people worked together, and acted locally and globally, to prevent the extinction of many species and ecosystems. Examples of successful conservation efforts are described throughout this textbook.

1.1. Conservation Biology is Still Evolving

As a distinct scientific field, conservation biology is an integrated, multidisciplinary subject that developed in response to the challenge of preserving populations, species, ecosystems, and biological interactions. The main aim of conservation biology is to ensure the long-term preservation of biodiversity. To achieve its aim, conservation biology has set three goals:

To document Earth’s biological diversity.To investigate how humans influence species, evolution, and ecosystem processes.To investigate practical approaches to protect and restore biological communities, maintain genetic diversity, and prevent the extinction of species.

The first two goals describe typical scientific research investigating objective facts. The third goal, however, is a part of what makes conservation biology a normative discipline; that is, conservation biology incorporates human values, not just facts, to understand and achieve its value-laden goals (Lindenmayer and Hunter, 2010). In this sense, conservation biology is related to environmentalism, in which people aim to protect the natural environment for its own sake (see Section 4.3.2). However, conservation biology is at its core a scientific discipline; it is founded on scientific principles. This is not to say you must be a scientist to practice conservation biology; there are many people who are not scientists who apply the principles of conservation biology in their professional and personal lives.

The emergence of conservation biology as a distinct scientific field in the 1970s has given rise to the formation of various formal societies representing the field in a united voice. Most notable among these is the Society for Conservation Biology (SCB, Figure 1.2), which is a non-profit international professional organisation with a mission to advance “the science and practice of conserving the Earth’s biological diversity”. To facilitate opportunities where like-minded people can share ideas locally, the SCB has regional branches, including an active Africa Section (http://conbio.org/groups/sections/africa) which hosts regular conferences. In addition to the SCB, a great number of other local, national, and regional conservation organisations also exist and act as mouthpieces for grassroots movements and as custodians of nature. Many of these groups focus on specific animals or local protected areas. Others organically adapt their missions and visions in response to a specific need or threat. For example, established in 1913 as an exchange forum between collectors of rare plants, the Botanical Society of South Africa now actively works toward protecting those rare plants in their natural habitats.

Figure 1.2The logo of the Society for Conservation Biology (SCB) has several layers of symbolism. Enclosed in the circle of life are ocean waves, representing change. The bird symbolises beauty, and the leaves (the bird’s wings) remind us of nature’s productivity. Image courtesy of SCB, all rights reserved.

In recent years, conservation practice has evolved from just a plan to save the environment to a vision that includes sustainable development and social justice.

Conservation biology also has a history of adapting to new challenges. The very first conservation activities, in Africa and beyond, were geared towards securing the rights to valuable natural resources for people in powerful positions, such as kings and tribal chiefs, enforced through a strictly adherence to cultural norms and customary laws (Section 2.2). But as a growing human population expanded its influence on the environment, and wildlife started to decline, earliest conservation models gradually shifted towards fortress conservation approaches (Wilshusen et al., 2002) which aimed to shield wildlife from people by setting aside protected areas where human activities were strictly controlled.

Today, however, as human populations are exploding, and consumption is increasing, even protected areas are increasingly unable to withstand the multitude of threats to biodiversity that ignore property boundaries and political borders. In response, fortress conservation approaches are beginning to make way for large-scale integrated activities that highlight the social and economic benefits of biodiversity conservation. To do this, new alliances are being formed and new agendas are being established, such as those that directly link human health with environmental health (Box 1.1). These integrated conservation philosophies that pursue strategies that benefit both humans and biodiversity show much promise because they focus on fundamental extinction drivers, and advocate for more inclusive sustainable development. In this way, the practice of conservation has evolved from just a plan to save the environment to a vision that accomplishes its goals through sustainable development and social justice.

Yet, as we consider how to best invest limited conservation resources, some difficult questions arise. With seemingly more work to be done than can be accomplished, should we let some species go extinct (Bottrill et al., 2008)? Which species? Who decides? How can we even dare to think that we can play god? Such questions predictably bring about strongly opinionated and emotional debate (Soulé, 2013 vs. Marvier, 2014; Tallis and Lubchenco, 2014). Given the successful track record of fortress conservation initiatives in preventing extinctions despite limited budgets (Young et al., 2014), as well as the promising progress of more complex people-centred initiatives (Pooley et al., 2014), it seems clear that conservation relies on some balance between these two conservation philosophies (Sodhi et al., 2011). Conservation biologists of tomorrow will be able to fine-tune the balance between these strategies by closely inspecting the successes and failures of our actions today.

Box 1.1 Conservation Through Public Health: A Case Study

Gladys Kalema-Zikusoka

Conservation Through Public Health,

Kampala, Uganda.

http://www.ctph.org

Conservation Through Public Health (CTPH) is a grassroots non-governmental organisation (NGO) and non-profit that promotes biodiversity conservation by enabling people, wildlife, and livestock to coexist. The organisation was founded in 2003 after fatal scabies skin disease outbreaks in mountain gorillas (Gorilla beringei beringei, EN) were traced to people living around Bwindi Impenetrable National Park, Uganda, who had limited access to basic health services (Kalema-Zikusoka et al., 2002). Since then, CTPH has contributed to conservation and sustainable development in Africa by improving human and animal health and welfare in and around protected areas.

One of the main goals of CTPH is to reduce disease transfer between humans and gorillas. We accomplish this through an integrated population, health, and environment (PHE) programme that was established in 2007 with funding from the US Agency for International Development (USAID). As a first step, piloted around Bwindi, CTPH held consultative meetings with local leaders, during which at least one Village Health and Conservation Team (VHCT) volunteer was selected from each village and two from each parish (consisting of 11 villages) to oversee distribution of family planning supplies. This initiative rapidly expanded into a sustainable social service delivery network that promotes family planning, hygiene, and sanitation. The network resulted in a 20% to 60% (national average is 30%) increase in new users to modern family planning, and a 10% to 60% increase in adoption of hand washing facilities at homes visited by VHCTs. VHCT volunteers also refer people suffering from infectious diseases and malnutrition to local health centres and promote more sustainable alternative livelihoods. Another group of community volunteers, the “Human and Gorilla Conflict Resolution” (HUGO) team, in turn collect gorilla faecal samples left on communal land to monitor their health (Figure 1.A), and visually monitor gorillas for clinical signs of disease inside and outside protected areas (Gaffikin and Kalema-Zikusoka, 2010). In the process, we have seen reduced disease incidences in the gorillas, reduced conflict between people and gorillas, and improved attitudes toward conservation. One unintended outcome has been increased gender equality: men are now more involved in family planning, and women are more involved in natural resource management.

Figure 1.AA park ranger from the Uganda Wildlife Authority teaching HUGO community volunteers how to collect faecal samples from gorilla night nests during a CTPH training workshop. Photograph by CTPH, CC BY 4.0.

Our experience in initiating and managing PHE programmes for the past 10 years has taught us several lessons. One of the most important lessons to ensure project sustainability is to regularly engage with local leaders and the government. The Uganda Wildlife Authority, Uganda’s Ministry of Health, and local health centres all attend CTPH meetings with VHCTs. Attendance by and representation of these groups not only informs them of our activities, but also provides a platform to inform or train the VHCTs in what they would like them to disseminate to the local communities.

We have also learnt that PHE-implementing partners and projects need to be well-suited to each other and each site; this remains true even though health needs are often the same, regardless of the location. For example, at Mount Elgon National Park in Uganda, we found that training VHCTs in reducing conflict with park management played a key role in changing community attitudes toward conservation. In contrast, at Virunga National Park, Democratic Republic of the Congo (DRC), we found that VHCTs needed to work more closely with local health centres to prevent disease transmission between people and gorillas, and to promote family planning in a largely Catholic country.

Lastly, we found that establishing income-generating projects for groups rather than individuals was key to sustaining VHCT networks and programme goals beyond donor funding cycles where we have had no volunteer dropouts in the first 10 years of initiating the PHE programme. These key components were accomplished by initiating livestock group enterprises and by encouraging VHCT volunteers to invest generated income into Village Saving and Loan Associations (see http://www.care.org/vsla).

1.2 The Role of Conservation Biologists

While there are a few extinctions that have only one cause, more generally, extinctions occur because several factors acted simultaneously and/or sequentially. Blaming a certain industry or specific group of people for an extinction (or other biodiversity loss) is thus simplistic, ineffective, and often counter-productive. Though challenging, a better approach would be to better understand how local, national, and international links led to those losses, and to find viable alternatives to prevent it from happening again. To succeed in this challenge, conservation biologists should strongly consider taking on one or more of several roles:

Conservation biologists should be curious. The world around us is full of natural wonders waiting to be discovered. These discoveries underpin conservation action, by allowing us to define all the different components of biodiversity, enabling us to better understand the needs of different species, and providing us with opportunities to celebrate our conservation successes.Conservation biologists must be good listeners. Sometimes, the only difference between attracting a new ally and making an enemy, or between developing a landscape and saving a species from extinction, is the way we communicate. Conservationists must be careful and respectful listeners, especially to opposing perspectives. Careful listening is particularly important in rural areas, where villagers often have practical concerns related to their daily contact with wildlife, such as staying safe and preventing crop damage and livestock loss. Quite often, those villagers may also have unique insights into wildlife ecology that could prove valuable in local conservation measures.Conservation biologists must be law-abiding citizens. Activities that involve wildlife and ecosystems are regulated by laws and regulations. These laws are important because ethical boundaries differ from person to person—activities acceptable to one group of people may be considered harmful by another. As conservation biologists, abiding by environmental laws is especially important if we want others to take those laws seriously.

Laws are important because ethical boundaries differ from person to person—activities acceptable to one person may be immensely harmful to another.

Conservation biologists should become effective communicators. They should be able to discuss the problems facing biodiversity in depth, as well as the consequences of losing biodiversity, to as broad a range of people as possible. Groups like hunters, community leaders and organisers, and church leaders may be interested in participating in conservation efforts once they recognise that their activities, health, and emotional well-being depend on conservation action. Conservation biologists could become politically active leaders, so that they can influence public opinion and policy. As a starting point, those interested in this role can join a conservation organisation to learn more about broader issues. They could also use their personal networks to form alliances with lawyers, citizen groups, and politicians. Conservation biologists could become pro-active land managers. Those taking on this task must be willing to walk on the land and go out on the water to find out what is really happening. They should also talk with local people to communicate their knowledge to others in ways that are clear and easily understood.Above all, a conservation biologist must be honest. To encourage effective action, both from the public and through policy, conservationists must present arguments backed by reliable evidence. To do otherwise, conservation biologists could lose credibility, which would very likely delay or even compromise conservation efforts.

It is worth taking a moment to distinguish between two important pillars of conservation action, namely conservation advocacy and conservation science. Conservation advocacy describes the roles that conservation biologists adopt to guide social, political, and economical systems towards a personally-preferred outcome—adopting environmentally-friendly practices; incorporating these activities makes conservation biology a normative discipline. Conservation science, in contrast, describes activities that conservation biologists undertake to generate knowledge, like objectively describing biodiversity and measuring biodiversity’s response to stressors and safeguards. While conservation advocacy and conservation science often support and inform each other as to the next steps required for “doing conservation”, it is important to distinguish between these two pillars to ensure that policy makers and other stakeholders in the environment understand when we advocate for personal preferences and when we offer objective findings (Rykiel, 2001; Lackey, 2007; Nelson and Vucetich, 2009). The next section will further expand on the importance of science in conservation biology.

1.3 The Value of Scientific Methods

The field of conservation biology applies scientific methods to achieving its goals. Like the medical sciences, which apply principles from physiology, anatomy, and genetics to problems of human health, conservation biologists solve biodiversity problems using principles from fields, such as mathematics, veterinary medicine, social sciences, and several natural sciences (Figure 1.3). Conservation biology differs from these and other component disciplines in that its primary goal is the long-term preservation of biodiversity. Unlike many other scientific fields, conservation biology can also be described as a crisis discipline (Soulé, 1985; Kareiva and Marvier, 2012). That is, conservation biologists are often required to take creative steps to respond to imminent threats, typically without a complete knowledge of the systems requiring attention. Conservation scientists must also articulate long-term visions for conservation beyond solving immediate problems.

Figure 1.3Conservation biology draws from many other sciences to protect biodiversity. It is closely related to natural resource management, which aims to manage biodiversity primarily for the benefit of humans. Integrated conservation and development projects (ICDP) are projects that manage nature for the benefit of both humans and biodiversity. After Kareiva and Marvier, 2012; Temple, 1991, CC BY 4.0.

To be effective, conservation biologists must demonstrate the relevance of their findings to a range of stakeholders. To be successful in this task, the importance of sound scientific principles cannot be over-emphasised. Nature is a complex network of many interdependent connections and feedback loops. Science is underpinned by principles that provide conservationists the necessary quantitative and qualitative tools to better measure and control for all these different aspects of biodiversity. Such measurements allow us to gain a better understanding of complex natural systems, and the consequences of human activities. Reliable, unbiased data obtained from sound and transparent scientific methods also facilitate policy making that is too often based on value judgments by non-experts who must balance many needs and different sources of information (Ntshotsho et al., 2015).

One of the cornerstones of modern science is to identify a hypothesis (a proposed explanation for a specific observation) to evaluate. The best hypotheses, often expressed as goals or objectives, are usually those that are SMART:

Setting specific, measurable, realistic and timebound goals and objectives is essential for effective conservation.

Specific: not overly general;

Measurable: has both units and a method of measurement;

Attainable: realistic to achieve;

Relevant: related to what needs to be accomplished;

Time-bound: achievable within a specific timeframe.

Identifying SMART goals and objectives is an essential aspect of conservation biology. Without such benchmarks, practitioners cannot know whether their tasks were successful, or when management actions should be adjusted to achieve success. While this may seem obvious, many previous conservation projects have failed because biologists neglected to set SMART goals and objectives (Tear et al., 2005). While lofty, “We’re going to save all species” is not a SMART conservation goal because it is overly general, hard to measure, unrealistic, and not time bound. In contrast, “We want to protect 25% of our country’s wetlands within the next 10 years” is a SMART goal because it sets a very clear and measurable objective. In general, it is wise to set smaller short-term (e.g. quarterly), and medium-term (e.g. annual) goals as one works towards long-term (e.g. 5–10 years) objectives; this allows one to constantly assess progress, which in turn provides opportunities for celebrations and strategic adjustments as and when needed.

Too many conservation decisions are based on biased anecdotes, personal intuition, and myths.

Another scientific standard that conservation practitioners must adopt at a larger scale is the transfer of knowledge gained from unique and specialist experiences. Conservation activities are too often hampered by the lack of guidance from credible and available sources. This forces conservationist managers to base important decisions on biased anecdotes, personal intuition, and even myths (Sutherland et al., 2004). Successful conservation actions on the other hand often rely on results and guidelines that were disseminated to the broader community by practitioners who faced similar challenges earlier. To maximise this learning from each other’s successes and mistakes, it is crucial for conservation scientists and managers to make every effort to ensure knowledge transfer, by carefully tracking their activities, and publishing their results and experiences in scientific journals and reports.

Public outreach builds on the public’s existing connection to nature and helps them better understand the value of local biodiversity.

Biodiversity conservation, however, is not accomplished by simply setting SMART goals, measuring outcomes, and publishing results in scientific journals and reports. It is also important for conservation biologists to engage in public outreach activities, during which they can build on the public’s existing connection to nature, help them better understand the value of biodiversity in their local area, and enable them to actively contribute in conservation projects. When interacting with the public, conservation practitioners must be sensitive to the complicated emotions and diverging interests of different groups of people (Milfont et al., 2017), especially vulnerable peoples who may be negatively impacted—hopefully only in the short term—by conservation actions. This requires a sense of emotional awareness, because the words we choose matter when we encourage others to care for and reduce their impact on nature. Equally important, conservation biologists, as with any field of science, should be sceptical of their results. The process of generating data is not equal to generating facts, because data can be fraught with bias, imprecision, and uncertainty. This is perhaps even more important when sharing findings with lay people, as scientists have rigorous training in understanding uncertainty and connecting cause and effect. Putting scientific findings in context with adequate and clear explanation is a challenge to all scientists, but it is necessary, especially when partnering with conservationists not specifically trained as scientists.

1.4 Environmental Ethics

Most human societies today aim to protect biodiversity through rules and regulations (Chapter 12). An alternative approach is to change the fundamental materialistic values of modern society to values that prioritise genuine and lasting human well-being. This is the goal of environmental ethics, a discipline within philosophy that emphasises the ethical values of biodiversity. The foundation of environmental ethics lies in the philosophical principle that every organism of Earth has a right to exist, regardless of its usefulness to humans, so any action that negatively impacts biodiversity would be considered unethical.

Because human quality of life is intricately linked to the ability of the natural world to prosper (Chapter 4), the ethical arguments for biodiversity conservation hold even for people who value only human life. Or, in other words, respect for human life—even our instincts for self-preservation—should compel us to preserve biodiversity. In contrast, if we neglect our assumed responsibility to act as guardians of life on Earth, future generations will suffer with a lower quality of life. We can already see signs of this today: as species are lost and natural ecosystems replaced with sprawling cities, children are increasingly deprived of the wonderful experience of seeing a ‘new’ animal (Figure 1.4) or pretty flower. We can imagine that we are borrowing Earth from future generations, and that it is our responsibility to ensure that they receive it in good condition.

Figure 1.4A ragged-tooth shark (Carcharias taurus, VU) fascinates two kids at the Two Oceans Aquarium, South Africa. People—and especially kids— enjoy seeing wildlife, as shown by the increased popularity of protected areas, zoos, and other institutions where biodiversity can be seen. Photograph by Karen Schermbrucker, courtesy of Two Oceans Aquarium, CC BY 4.0.

Because of this close link between nature and human well-being, the concept of nature preservation has permeated through the value systems of most human cultures, philosophies, and religions throughout history. This is especially relevant in Africa, where most (if not all) traditional societies have a deep connection with nature that is woven into their spiritual beliefs and customs (Figure 1.5). Our responsibility to protect animals is also explicitly described in Jewish, Christian, and Islamic traditions. Other major religions, including Hinduism, Buddhism, and Taoism also strongly support the preservation of non-human life. In light of accelerated biodiversity losses, faith-based groups have recently started playing a more active role in conservation, particularly among urbanised people. They do this by informing adherents that it is wrong to allow the destruction of nature, and that such destructive activities can have negative consequences for all people on Earth. These links between faith-based organisations and conservation have given rise to consortiums such as the Forum on Religion and Ecology (http://fore.yale.edu), the Alliance of Religions and Conservation (http://www.arcworld.org), and the SCB’s Religion and Conservation Working Group (https://twitter.com/ReligionConBio), as well as the emerging field of spiritual ecology (Vaughan-Lee, 2016).

Figure 1.5The spiritual connection between people and nature features strongly in ancient rock art made by Bushmen (also known as San, or First People) of Southern Africa, believed to be the oldest human population on Earth. Pictured here is a common eland (Tragelaphus oryx, LC), drawn by shamans to open the portals of the spiritual world. Photograph by Alan Manson, https://commons.wikimedia.org/wiki/File:Shaman_at_work_Game_Pass_2004_0522_121946AA.jpg, CC BY 4.0.

Environmental ethics has strong links to the environmental justice movement and has recently established strong ties to the social justice movement. Some of the most exciting developments in this direction involve initiatives that combine protection activities with community upliftment programmes that improve the well-being of local peoples (Box 1.2; see also Section 14.3). These developments have shown that when poor and marginalised people are empowered to protect the environment, they may act as strong local guardians of forests, coastal areas, and other ecosystems that may have been destroyed otherwise.

Box 1.2 The Okapi Wildlife Reserve: Protecting Nature and Providing for People

Rosmarie Ruf & Marcel Enckoto

Okapi Conservation Project,

Okapi Wildlife Reserve,

Epulu, DRC.

https://www.okapiconservation.org

The Okapi Wildlife Reserve, a World Heritage Site in peril, is located within the dense, tropical Ituri Forest in north-eastern DRC. The reserve was created to protect the okapi (Okapia johnstoni, EN) (Stephenson and Newby, 1997). Researchers estimate that there are between 10,000 and 25,000 okapi (Figure 1.B) left in the wild, but with populations appearing to have declined by more than 50% over the last 15 years (Kümpel et al., 2015). The reserve also protects charismatic species like forest elephants (Loxodonta cyclotis), 14 species of primates, including chimpanzees (Pan troglodytes, EN), leopards, forest buffalo (Syncerus caffer nanus), and bongo antelope (Tragelaphus eurycerus, NT).

Figure 1.BThe okapi, a relative of the giraffe, at ZooPark de Beauval, France. This species, one of the DRC’s natural treasures, survives only because of the dedication of a devoted group of conservation biologists. Photograph by Daniel Jolivet, https://www.flickr.com/photos/sybarite48/7973333500, CC BY 2.0.

Since 1987, the Okapi Conservation Project has partnered with the Congo Institute for Conservation of Nature, the government agency responsible for the Okapi Wildlife Reserve’s management, to provide financial and technical support for the operation of the reserve and preservation of the surrounding Ituri Forest. The project is partially managed and funded by Wildlife Conservation Global, a non-profit NGO based in Florida, USA.