Chapter 1

9 downloads 0 Views 10MB Size Report
PA includes all the sites that are always selected for conservation regardless of the RS method .... In 2012 Cameroon counted 25 national parks (NPs), 3 of which were .... domination that has accelerated the degradation of species and pushed forest dwellers ...... compromising their well-being and those of their children.
Complimentary Contributor Copy

Complimentary Contributor Copy

ENVIRONMENTAL SCIENCE, ENGINEERING AND TECHNOLOGY

NATIONAL PARKS SUSTAINABLE DEVELOPMENT, CONSERVATION STRATEGIES AND ENVIRONMENTAL IMPACTS

No part of this digital document may be reproduced, stored in a retrieval system or transmitted in any form or by any means. The publisher has taken reasonable care in the preparation of this digital document, but makes no expressed or implied warranty of any kind and assumes no responsibility for any errors or omissions. No liability is assumed for incidental or consequential damages in connection with or arising out of information contained herein. This digital document is sold with the clear understanding that the publisher is not engaged in rendering legal, medical or any other professional services.

Complimentary Contributor Copy

ENVIRONMENTAL SCIENCE, ENGINEERING AND TECHNOLOGY Additional books in this series can be found on Nova’s website under the Series tab.

Additional e-books in this series can be found on Nova’s website under the e-book tab.

Complimentary Contributor Copy

ENVIRONMENTAL SCIENCE, ENGINEERING AND TECHNOLOGY

NATIONAL PARKS SUSTAINABLE DEVELOPMENT, CONSERVATION STRATEGIES AND ENVIRONMENTAL IMPACTS

JOHNSON B. SMITH EDITOR

New York

Complimentary Contributor Copy

Copyright © 2013 by Nova Science Publishers, Inc. All rights reserved. No part of this book may be reproduced, stored in a retrieval system or transmitted in any form or by any means: electronic, electrostatic, magnetic, tape, mechanical photocopying, recording or otherwise without the written permission of the Publisher. For permission to use material from this book please contact us: Telephone 631-231-7269; Fax 631-231-8175 Web Site: http://www.novapublishers.com NOTICE TO THE READER The Publisher has taken reasonable care in the preparation of this book, but makes no expressed or implied warranty of any kind and assumes no responsibility for any errors or omissions. No liability is assumed for incidental or consequential damages in connection with or arising out of information contained in this book. The Publisher shall not be liable for any special, consequential, or exemplary damages resulting, in whole or in part, from the readers’ use of, or reliance upon, this material. Any parts of this book based on government reports are so indicated and copyright is claimed for those parts to the extent applicable to compilations of such works. Independent verification should be sought for any data, advice or recommendations contained in this book. In addition, no responsibility is assumed by the publisher for any injury and/or damage to persons or property arising from any methods, products, instructions, ideas or otherwise contained in this publication. This publication is designed to provide accurate and authoritative information with regard to the subject matter covered herein. It is sold with the clear understanding that the Publisher is not engaged in rendering legal or any other professional services. If legal or any other expert assistance is required, the services of a competent person should be sought. FROM A DECLARATION OF PARTICIPANTS JOINTLY ADOPTED BY A COMMITTEE OF THE AMERICAN BAR ASSOCIATION AND A COMMITTEE OF PUBLISHERS. Additional color graphics may be available in the e-book version of this book.

Library of Congress Cataloging-in-Publication Data ISBN:  (eBook)

Published by Nova Science Publishers, Inc. † New York

Complimentary Contributor Copy

CONTENTS Preface Chapter 1

Chapter 2

Chapter 3

Chapter 4

vii The Tragedy of the Governmentality of Nature: The Case of National Parks in Cameroon Ngambouk Vitalis Pemunta, and Ogem Pascal Mbu-Arrey Linking Protected Area Conservation with Poverty Alleviation in Uganda: Integrated Conservation and Development at Bwindi Impenetrable National Park J. Baker, R. Bitariho, A. Gordon-Maclean, P. Kasoma, D. Roe, D. Sheil, M. Twinamatsiko, G. Tumushabe, M. van Heist and M. Weiland Community-Based Conservation: An Institutional Approach to Assessing Biodiversity Conservation Efforts at Bardia National Park in Nepal Shova Thapa Karki Priority Areas Integration (PAI) Method: A Tool to Facilitate Biodiversity Conservation? Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

Chapter 5

Communal Game Farming: A Sustainable Land Use Option? V. Kakembo, P. Mamfengu and G. Kerley

Chapter 6

Land As Sustenance and Sanctuary: Settlement History and Resource Use in and around Utah’s Grand Staircase-Escalante National Monument Robert J. Lilieholm and Marietta Eaton

Chapter 7

Assessing the Competitiveness of Protected Areas: The Competitive Positioning of the Peneda-Gerês National Park Maria João Carneiro and Carlos Costa

Complimentary Contributor Copy

1

47

105

129

155

179

201

vi Chapter 8

Contents The Doñana National Park (SW Spain): From the Sea to the Future F. Ruiz, M. Pozo, M. Olías, M. C. Núñez, M. Abad, M. I. Carretero, J. Rodríguez Vidal, L. M. Cáceres, J. Prenda, E. M. Castellanos, C. J. Luque, A. Menor, E. Font, A. Toscano and E. X. García

Index

219

233

Complimentary Contributor Copy

PREFACE In this book, the authors present current research in the study of sustainable development, conservation strategies and the environmental impacts of National Parks. Topics discussed include the case of National Parks in Cameroon; linking protected area conservation with poverty alleviation at Bwindi Impenetrable National Park, Uganda; biodiversity conservation efforts at Bardia National Park in Nepal; Priority Areas Integration (PAI) methods as a tool to facilitate biodiversity conservation; communal game farming; settlement history and resource use in Utah's Grand Staircase-Escalante National Monument; the competitive positioning of the Peneda-Geres National Park in Portugal; and the Donana National Park in Spain. Chapter 1 – Cameroon’s rich humid tropical rainforest is degrading at an alarming rate. The government has deployed several governmentality regimes including fortress conservation through the establishment and gazzetment of national parks and protected areas to arrest this disastrous situation. On the basis of two case studies, this chapter explores the evolution of the governance of national parks and the institutional obstacles that make protected areas a parody in Cameroon. This essay argues that the creation and implementation of national parks as artifacts and processes illustrate the nature and extent of global governmentality upon these regions, their people and their natural resources. While the protection of national parks and other reserves ensures that biological diversity is secured and also allows natural processes such as evolution to continue undisturbed by human influence, the exclusionary nature of protected area management regimes however, highlight the chasm between law as a governmental rationale which often fails to recognize traditional usage rights and instead tends to create conflicts between protected areas and local communities. This scenario is compounded by the acute underdevelopment that characterizes protected areas where top-down development is the norm. The paper calls for a middle ground that balances both local and national, as well as national and international interests as the basis for achieving sustainable people-centered development and conservation. Reaching this goal will entail a radical policy shift from the present paramilitary approach of outright protectionism, which is a recipe for disaster to a system that takes account of the interests of local people while still conserving biodiversity as well as the re-institutionalization of customary land tenure norms with new legal provisions. Furthermore, sustainable national park management practices are needed. This could include the creation of a Conservation Development Authority with the mandate to initiate private sector partnership and also encourage community development and participation in the effective management and control of protected areas. The mandate of this body must incorporate active community participation in decision-making and planning for the

Complimentary Contributor Copy

viii

Johnson B. Smith

sustainable use of ecosystem services and development of ethno-tourism. If trends in rural emigration and depopulation are to be halted, then national parks must to be protected in line with sustainability principles. Chapter 2 – Gaining the support of local communities for conservation and resolving local conflict issues are priorities for managers of national parks. Conflict can be defined as the expression of divergent interests between resource-poor households neighbouring a national park and the national and international actors concerned with conservation of biological diversity. Conflict can arise when access to natural resources is prohibited or from human-wildlife conflict. In addition to gaining the support of local communities, managers must also protect endangered wildlife and ecosystems from activities that threaten the conservation status, particularly unauthorized resource use. However achieving the balance of improving relations with local communities while enforcing conservation law can be a significant challenge, particularly at national parks surrounded by high populations of rural communities whose livelihoods depend on the natural resource base. Chapter 3 – Limited success in community-based conservation has challenged the role of institutions and design of incentive-based approaches. Earlier research has indicated the importance of a supportive institutional framework for the provision of effective incentives. This paper examines this issue for biodiversity conservation at Bardia National Park in Nepal. Derived from a wide survey of literature, seven criteria to assess institutional performance at Bardia were identified. Issues such as illegal resource extraction, park-people conflict, and deterioration of trust in NGOs indicated institutional failure. On the other hand, the park management has supported local communities by forming financial savings groups, and establishing a buffer zone management council to build social capital. This has resulted in increased collective action, greater trust within communities, enhanced citizen trust in the park management and more positive attitudes towards biodiversity conservation. Findings presented here indicate that even though the initial design of conservation initiatives often includes some of the characteristics necessary for success, these tend to be lost in the implementation process, with individuals in positions of responsibility shaping the process to suit their interests. Currently, the State acts as a mediator between the local communities and NGOs. This research suggests that instead, the communities themselves should be given a central position between the State and the NGOs, with all programmes channelled through them. By enhancing technical capabilities, programme monitoring, and building of institutions through trust, cooperation and collective action, such a rearrangement of roles would contribute not only to an effective design of conservation initiatives, but also to their effective implementation. Chapter 4 – In biodiversity conservation planning a keystone process is the selection of the proper priority area and for that reason many diverse reserve selection (RS) methods have been developed. Usually depending on the chosen method a unique optimum solution is estimated based on the calculation of specific ecological criteria, indicators and algorithms. The objective of this chapter is the design of a method (called PAI) that integrates priority areas estimated from different RS methods. From this integration, through a simple process of computational steps, equivalent areas of priority are estimated which generate alternative conservation scenarios. The total area of conservation concern is separated into three major sub-priority areas (PA) called areas of high (H), medium (M) and low (L) priority. The highPA includes all the sites that are always selected for conservation regardless of the RS method used, while the low-PA includes all the sites that are always rejected, respectively. Medium-

Complimentary Contributor Copy

Preface

ix

PA includes all the sites that might be selected or rejected due to the RS method applied and their combination provides alternative solutions of priority areas. The insular areas of the Mediterranean basin were selected to implement the PAI method on a regional scale. In particular, in the South-East Aegean region a group of islands (Crete, Rhodes, Symi and Tilos) were studied as typical examples of combining high values of biodiversity and strong economic development, mainly tourism. The results for all the islands show significant levels of concordance between potential sites. Spatial distribution of the various biodiversity values increases proportionally according to the island’s size, whereas in smaller islands the values from the different RS methods overlap. Alternative scenarios emerge for every potential size of the conservation area, providing the ability to implement socio-economic criteria into the process if needed, especially in areas undergoing intense development such as Crete and Rhodes. A generalized PA graph of the PAI method that analyses and presents a general spatial distribution of the conservation values of a conservation concern area is proposed. This unified graph, unique for each area of interest, visualizes the congruencies and variances of these values and based on their integration, produces conservation alternatives rather than defining an optimum solution. Such scenarios could facilitate the planners to design and implement more effective strategies in conservation planning. Chapter 5 – Severe land degradation in many communal villages of South Africa has rendered land use forms such as cultivation and livestock farming unviable and unsustainable, making it difficult to earn a land based livelihood. Sustainable communal land management is achievable through the provision of more land use options which are more environmentally friendly. Against the backdrop of land degradation, which is coupled with high poverty levels in the communal villages, game farming has been recommended by ecologists as an ecologically, economically and socially sustainable form of land use. Wildlife conservation in South Africa occurs predominantly at two levels, namely, state owned Parks (National, Provincial and Municipal) and Private Game Reserves. The involvement of local communities has been identified by many reserchers as the key to meaningful and effective wildlife conservation. In the Eastern Cape Province, the Tyefu Community Reserve project has been recommended by the Subtropical Thicket Ecosystem Project (STEP), backed by the Department of Environment Affairs and Tourism (DEAT) as a sustainable land use option. Enterprises of this nature have the potential to alleviate poverty whilst conserving biodiversity. However, the following environmental questions arise before such an undertaking is embarked upon: Is the proposed area for the reserve a suitable habitat for wildlife? What kind of wildlife species can be supported, and what is the carrying capacity for the proposed community reserve? How feasible would game farming be in a communal setting against the backdrop of land tenure and land-use constraints? The following aspects are analysed in this case study: the abundance and condition of vegetation, terrain parameters, potential wildlife species and the carrying capacity of the recommended reserve. These should provide the basis for discussion of communal game farming as a land use option in other parts of the country. Chapter 6 – President Bill Clinton designated the 1.9-million-acre Grand StaircaseEscalante National Monument (GSENM or Monument) in September of 1996. While the Monument’s creation was largely intended to protect the region’s scientific and historic values, the growing economic importance of recreation and tourism was an important though less-noted factor. In this respect, the establishment of GSENM represents the culmination of protection efforts and counter-efforts in the region that date back to the 1930s – when parts of

Complimentary Contributor Copy

x

Johnson B. Smith

today’s Monument were first considered for federal protection within the USDI National Park Service. The prolonged struggle to preserve the area’s remoteness has sparked a string of controversies – a feature of the land and culture that will likely endure. This chapter describes the social, economic and cultural history of the people that settled in and around GSENM in an effort to provide context for past, current, and future debates surrounding the Monument’s creation and management. The lessons learned here provide useful insights for other regions of the globe seeking to balance resource protection and changing social demands. Chapter 7 – Protected areas stand out due to the importance of specific nature and biodiversity features they comprise. However, several researchers highlight that the competitiveness and consequent sustainability of tourism destinations do not only depend on a specific resource of the destination, but on the whole basis of resources of the destination and on the way this set of resources is managed. Competitive positioning studies became an important tool for evaluating the competitiveness of destinations, since they enable to understand how potential visitors evaluate the destination comparing to potential competitors. Therefore, this kind of studies provides important contributions for improving the management and competitiveness of destinations. Despite the importance of this type of research, positioning studies on protected areas are rare. The present chapter aims to: highlight the importance of competitive positioning studies of protected areas; present a methodology that may be used for assessing the competitive positioning of protected areas; and to assess the competitive positioning of a specific national park located in Portugal - the Gerês National Park. It is also aimed to provide contributions for a more efficient management of the Gerês Park, in particular, and of protected areas, in general. A total of 1115 questionnaires directed to visitors of Gerês were obtained. Results highlight the importance of assessing the positioning of protected areas considering the performance of these areas regarding the following features: tourism attractions; facilities; protected areas’ ability to satisfy motivations; and the constraints visitors feel for visiting these areas. The empirical research undertaken reveals that Gerês holds competitive advantages in relation to competing destinations in several domains, such as on natural attractions, peacefulness and on providing opportunities for escaping and relaxing. However, the study also highlights the need to further explore the potential of protected areas, by complementing the important natural supply of these areas with, for example, an attractive and well managed set of cultural resources and interesting opportunities for socializing. The provision of information about the protected area reveals to be a good strategy for improving the attractiveness of protected areas, as well as for reducing constraints for visiting these areas. Chapter 8 – The Doñana National Park is one of the most important National Parks in Spain. The present-day scenario is the final result of a long geological evolution (~6 Myr), with the ´creation´ of a deep aquifer and numerous geomorphological features that enhance the biodiversity. This socio-ecological system presents numerous threats (invasive species, decrease of aquifer discharges, tourism resorts, sea level rise) and an adequate sustainable development is necessary in order to preserve this special system.

Complimentary Contributor Copy

In: National Parks Editor: Johnson B. Smith

ISBN: 978-1-62618-934-8 © 2013 Nova Science Publishers, Inc.

Chapter 1

THE TRAGEDY OF THE GOVERNMENTALITY OF NATURE: THE CASE OF NATIONAL PARKS IN CAMEROON Ngambouk Vitalis Pemunta1, and Ogem Pascal Mbu-Arrey2,† 1

Linnaeus University Centre for Concurrences in Colonial and Postcolonial Studies, Linnéuniversitetet, Växjö, Sweden 2 Department of Sociology and Anthropology, Central European University, Budapest, Hungary

ABSTRACT Cameroon’s rich humid tropical rainforest is degrading at an alarming rate. The government has deployed several governmentality regimes including fortress conservation through the establishment and gazzetment of national parks and protected areas to arrest this disastrous situation. On the basis of two case studies, this chapter explores the evolution of the governance of national parks and the institutional obstacles that make protected areas a parody in Cameroon. This essay argues that the creation and implementation of national parks as artifacts and processes illustrate the nature and extent of global governmentality upon these regions, their people and their natural resources. While the protection of national parks and other reserves ensures that biological diversity is secured and also allows natural processes such as evolution to continue undisturbed by human influence, the exclusionary nature of protected area management regimes however, highlight the chasm between law as a governmental rationale which often fails to recognize traditional usage rights and instead tends to create conflicts between protected areas and local communities. This scenario is compounded by the acute underdevelopment that characterizes protected areas where top-down development is the norm. 

Ngambouk Vitalis Pemunta: Linnaeus University Centre for Concurrences in Colonial and Postcolonial Studies, Linnéuniversitetet, SE-351 95 Växjö, Sweden. † Ogem Pascal Mbu-Arrey: Department of Sociology and Anthropology, Central European University, Zrinyi utca 14,1051, Budapest, Hungary.

Complimentary Contributor Copy

2

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey The paper calls for a middle ground that balances both local and national, as well as national and international interests as the basis for achieving sustainable people-centered development and conservation. Reaching this goal will entail a radical policy shift from the present paramilitary approach of outright protectionism, which is a recipe for disaster to a system that takes account of the interests of local people while still conserving biodiversity as well as the re-institutionalization of customary land tenure norms with new legal provisions. Furthermore, sustainable national park management practices are needed. This could include the creation of a Conservation Development Authority with the mandate to initiate private sector partnership and also encourage community development and participation in the effective management and control of protected areas. The mandate of this body must incorporate active community participation in decision-making and planning for the sustainable use of ecosystem services and development of ethno-tourism. If trends in rural emigration and depopulation are to be halted, then national parks must to be protected in line with sustainability principles.

ABBREVIATIONS ACP AIDS CBNRM CAR CCS CFA DBR DFC ECOFAC EU FAO GIS GTZ GoC Ha ICDPs IIED IUCN KNP MINEF MINFOF ND NFAP NFE NGO NPFE NPS ODA ODI

African-Carribean and Pacific Group of States Acquired Immune Deficiency Syndrome Community-based Natural Resource Management Central African Republic Community Conservation Strategy Common Francs in Africa Dja Biosphere Reserve Direction de la Faune et de la Chasse Central African Forestry Ecosystems European Union Food and Agriculture Organization Geographic Information System the German Technical Cooperation Government of Cameroon hectares International Conservation and Development Programmes International Institute for Environment and Development the International Union for Conservation of Nature Korup National Park Ministry of Environment and Forestry Ministry of Forestry and Wildlife No Date National Forestry Action Plan National Forest Estate Non-Governmental Organization Non-Permanent Forest Estate National Parks Organisation for Development Assistance Oversea Development Institute

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature PAs PFE SG-SOC SFM TNCs UNDP UNO UNESCO UK US USD WRM WWF ZIC ZICGC

3

Protected Areas Permanent Forest Estate SG Sustainable Oils Cameroon Sustainable Forestry Management Transnational Corporations United Nations Development Programme United Nations Organisation United National Educational Scientific and Cultural Organisation United Kingdom United States United State Dollar World Rainforest Movement Worldwide Fund for Nature Hunting zone (Zone d’intérêt cynégétique) Hunting zone under community-based management (Zone d’intérêt cynégétique à gestion communautaire)

INTRODUCTION Cameroon has an estimated 22 million hectares of forest resources containing about 1.5 billion m3 of timber of its total land surface of 475,000 square kilometers. Unfortunately, about 200,000 hectares of these natural resources are lost annually while over 40 species of wildlife are at the verge of extinction due to unsustainable exploitation (WRM, 1999, Fonjong, 2006:664). As of 2005, 14% of Cameroon’s territory had been earmarked as protected areas (PAs). In 2012 Cameroon counted 25 national parks (NPs), 3 of which were UNESCO-MAB biosphere reserves (Benoue: 180,000, Dja: 500,000 and Waza: 170,000 hectares) and one, a UNESCO heritage site, 5 wildlife reserves, 3 wildlife sanctuaries, 3 wetlands of international importance, 16 forest reserves and PAs, 9 cloud (mountain) forest sites (United Nations List of Protected Areas,1 Kimbi, 2012, Mertens et al., 2012, MINFOF, 2011, UNESCO, 2008). From 2006 to 2011, land allocated to PAs increased by 8%, to 7.4 million hectares (16% of total national land area and 43 % of the total National Forestry Estate (NFE). As of June 2011, there were a total of 86 PAs (including 52 hunting zones) covering about 45% of the Permanent Forest Estate (PFE)—an area increase of 8% (554,217 ha) since 2006. The main driver behind this increment was the creation of 10 new NPs- many of which resulted from the reclassification of forest reserves2 (Mertens et al., 2012:29). National parks (NPs) are created as pristine wilderness reserves, isolated from adverse human impacts and simultaneously making parks accessible to the public (Lowry 1994,

1 2

http://www.unep-wcmc.org/un-list-of-protected-areas_269.html. Last accessed Januray 16, 2013. These include the Kagwene, Mefou, Ebo, Ndongere, Mount Cameroon, Takamanda, Tchabal Mbabo, Kom, Ma Mbed Mbed, and Deng Deng NPs. Of these, Takamanda, Ndongere, Ebo, and Deng Deng NPs were formerly forest reserves, and the Mount Cameroon NP now includes the former Bomboko and Etinde Forest Reserves. Consequently, a total of six former forest reserves were converted into new national park land between 2007 and 2009. The shift from forest reserve, which is technically a designated production forest, to a national park—an integrally protected entity—constitutes a significant change in land use classification, since it is much easier to change a “reserve” to a “park” than vice versa (Mertens et al., 2012:29).

Complimentary Contributor Copy

4

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

Kopas 2008) such as for education and for tourism. Forests and trees simultaneously provide a wide range of socio-economic, cultural and environmental benefits to humankind. Table 1. Categories of Protected Areas in Cameroon Protected Areas3

No

National Parks 24 Wildlife Reserves 5 Sanctuaries 5 (Floral and faunal) Hunting zones 52 (ZIC/ZICGC) Total 86 Source: Mertens et al. (2012:17).

3,459,798 715,456

% National Forest Estate 20 4

% of National Total4 7 2

143,909

1

< 1/2

3,078,418

18

7

7,397,581

43

16

GIS Area (ha)

They deserve conservation and wise use for the benefit of rural people who are dependent on these forests and for the benefit of humanity in general. In Cameroon, policymakers have tried to resolve the conflict between ecological protection and support for the enjoyment of present and future generations of park visitors through the adoption of sustainable development principles such as through the integration of the economy and the environment and by adopting poverty alleviation policies that mitigate the environmental and social costs of development. Forestry policy is a central component of Cameroon’s poverty alleviation and national environment programme. This comprehensive document outlines government priority areas in resource management as well as acknowledges the role of women in conservation. Environmental protection has further been sustained through related international, regional and sub-regional treaties and conventions in the domain of natural resource management. The Brundtland Commission defined sustainable development as “development that meets the needs of the present without compromising the ability of future generations to meet their own needs” (United Nations Organisation, 1987). In other words, sustainable development must both satisfy human needs and at the same time include responsible use of society’s scarce resources. This lofty idea has been integrated into national legislation most notably through Cameroon’s 1994 Forestry Law and its subsequent decree of application. While the idea of creating NPs is salutary, it is fraught with contradictions and has instead led to the violation of the human rights of forest dwellers and to the intensification of environmental degradation. Through two case studies- the Dja Biosphere Reserve (DBR) straddled between the east and southern regions and the Korup National Park (KNP) in Southwest Cameroon, this chapter examines the meaning and practice of sustainable development. The purpose of this 3

Protected areas figures for 2009 and 2011 include the recently proposed national parks (Kom, Mefou, Ebo, Tchabal Mbabo, Ndongere, and Ma Mbed Mbed) and the Rumpi Hills Sanctuary, which are still pending official classification. Together, these proposed protected areas account for 609,221 ha (cf Merens et al. 2012). 4 Total land area of Cameroon is listed at between 466,326 km2 (de Wasseige et al., 2008) and 472,710 km2 (CIA Factbook, 2011). The total land area (468,305 km2 excluding maritime territory) in Table 1 is based on the Cameroon national boundary included in V3.0 (cf, Ibid).

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

5

chapter is to demonstrate that the creation of NPs as protected spaces is an exercise in domination that has accelerated the degradation of species and pushed forest dwellers deeper into poverty. According to the World Bank, forest deforestation in Cameroon increased from 0.4% between 1976 and 1980 to 0.6% annually between 1981 and 1996. It is further estimated that since 2000, the rate of deforestation stands at approximately 222,000 hectares- corresponding to a deforestation rate of 0.88% per year (Samuel and Urie, 2011). Commenting on the link between deforestation and projected annual population growth rate, the Food and Agricultural Organization (FAO) estimates that between 2010 and 2015 the rate of forest depletion in the country will be approximately between 300,000 hectares per year and 400,000 hectares annually between 2030 and 2040 (FAO, 1997, Samuel and Urie, 2011). Having lost nearly 2 million hectares of its historic close forest cover between 1980 and 1995, Cameroon now has the second highest deforestation rate in the Congo Basin after the Democratic Republic of Congo (Hutter, 2000). The opening of new lands for agriculture as well as logging activities and energy consumption is the main drivers of deforestation in sub-Sahara Africa (UNDP, 1995, Fonjong, 2006: 666). Table 2. Cameroon forest management 1993 forest cover Protected area 1999 concession area Protected area as a percentage of 1993 forest cover 1999 Concession area as a percentage of unprotected forest cover 1959–1999 cumulative concession area as a percentage of unprotected forest cover Source: Global Forest Watch (cf Fonjong, 2006:667).

22.8 million hectares 1.4 million hectares 7.7 million hectares 6.1% 36% 81%

Table 3. Cameroon: Trends in Natural Forest Cover (Deforestation), 1990-20105 FOREST COVER (excluding planted forests) (1000 ha) 1990 2000 24316 22116 ANNUAL CHANGE RATE (1000 ha) Negative number represents deforestation 1990-2000 2000-2005 -220 -220 ANNUAL CHANGE RATE (percent) Negative number represents deforestation 1990-2000 2000-2005 -0.9 -0.90

5

2005 20932

2010 19916

2005-2010 -220

2005-2010 -0.99

WWW. http://rainforests.mongabay.com/deforestation/2000/Cameroon.htm. Last accessed 16 January, 2013.

Complimentary Contributor Copy

6

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

About 18 million hectares of Cameroon’s original forest estate has been cleared down for agriculture and settlement6, with nearly two million hectares alone lost between the periods 1980 and 1995 (The Courier, 2002). Cameroon has thus lost almost half of its historic forest. Deforestation has continued to increase as depicted below. This chapter is based on a review of related literature and reliance on oral evidence and firsthand observation to reconstruct the social conditions of forest dwellers around the KNP and DBR in Cameroon. Individual in-depth interviews were conducted with different stakeholders- forest users and representatives of conservation organizations as well as the local public administrative authorities- in these areas in 2009 and 2010 respectively. During these interviews, themes such as forest exploitation, the activities of both governmental and Non-Governmental conservation agencies and the perception by the forest communities of the various strategies of external interventions in the management of their forest were explored. This was complemented by participant observation of forest use activities among various groups of forest users. The problems of these NPs are representative of problems and trends in the entire West and Central African regions-which increases the pertinence of this study to conservation and development beyond the local level. The aim is to identify the disjunctures between conservation and development that are rooted in institutional obstacles and the implementation of suggested recommendations to reduce conflict and challenges to sustainability.

FORTRESS CONSERVATION AS A GOVERNMENTAL RATIONALE: CONTINUITIES AND DISJUNCTURES This section provides a brief overview of the evolution of land tenure from the colonial to the postcolonial period in Cameroon in order to situate the institutional context within which PAs and NPs are established and to highlight the source of protracted conflicts between local communities, conservation NGOs and the government of Cameroon (GoC). Forest conservation as a regime of governmentailty began during the colonial era and was intertwined with the colonial notion of “private property”7 and spatial governmentality. The colonial government nationalized natural resources (Frost et al. 2006) and large tracts of land were set aside for conservation. This led to the displacement, marginalization and alienation of local people from traditional resource areas (Colchester 1994, Davenport and Usongo 1997). This trend continued after independence and the “commercial uses of natural resources” has remained “centralized, conditioned by government policies of the colonial and postcolonial eras” (Frost et al. 2006:16). The colonial regime transferred land ownership from traditional local authority to the state domain to enable the state to exploit African lands, labour, and resources. Resources such as wildlife were progressively placed under a central regulatory authority, with the resource user rights of local people alienated over time. This shift in tenure later became one of the key drivers of African nationalist movements seeking to recover entitlements to land 6 7

Global Forest Watch, n.d., http://www.globalforestwatch.org/english/#brief. Land ownership entailed that it was being put into productive use say through farming. This is reminiscent of the mis en valeur principle in Cameroonian land tenure law.

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

7

and resources (Roe et al. 2009:6) such as the Mau Mau uprising in Kenya. The European preoccupation with the commercial exploitation of natural resources such as timber, ebony, rubber, kola nuts and palm oil from natural forest, later led to large tracts of land being cleared to make way for the cultivation of cash crops-basically cocoa and oil palm (Martin, 1991:15). Although the colonial regime undertook large scale reafforestation by planting thousands of exotic and native timber species, they later came to the stark realization that a totally protectionist policy was more appropriate than the planting of trees (Martin, 1991:18). Alarmed by human-induced forest degradation through clearance for farming and to a lesser extent logging, the dominant orthodox view of overwhelming and rapid deforestation became the institutionalized policy framework within West African colonial administrations and policies (Fairhead and Leach, 1998:168). In Cameroon like in most West African French colonies, state intervention was premised on the fallacy that “primary” forests were not in use and were “seen primarily as forests in which the local people have never exercised any right... They are undoubtedly of the character of forest empty and without owners” (“vacantes et sans maître”) (Auberville, 1949: 100, Fairhead and Leach, 1998: 169). In other words, they were ownerless. Cameroon’s tripolar colonial heritage—German, British and French has further complicated the country’s land tenure systems. For example, the Germans converted collective or communal land to individual ownership under the assumption that all unoccupied land was vacant land without an owner and such land became “kronland”, i.e., German land (Logo et al. 2003:48). Similarly, the French 1938 Decree named all unoccupied land as “Terre Vacant et sans Maître” (land without owner). In this regard, native land rights were trampled upon as many collective (customary) land rights were transformed and registered as titled land with a new owner-the colonial state and later, the postcolonial state. While the British recognized natives as the sole owners of vacant land, such land was under the control of the state (see the Native Rights Ordinance of Northern Nigeria No. 1 of 1916 and Southern Cameroon No. 1 of 1927 (Logo et al. 2003:48). This has led to ongoing struggles over ownership of land and land resources between the State, the community and the individual. In former British Cameroon-(today’s Northwestern and Southwestern regions), rights were vested in the hands of “Native Authorities” under the British colonial policy of indirect rule. The focus was on the reservation of large tracts of native lands carved out as “virgin forest”, as land reserves- with the covert aim of logging them or appropriating them for agriculture (Mayers and Bass, 1998:276). A typical case in point was the appropriation of Bakweri land for the establishment of the Cameroon Development Corporation-an agro-industrial complex that has been the bone of contention and litigation between the GoC and the Bakweriland Claims Committee (for a detailed historico-anthropological analyses of the claims and counterclaims of both parties see Pemunta and Fonmboh, 2010). Reserves were also created for the protection of watershed, as shelter-belts (Martin, 1991: 18) and for the maintenance of climatic conditions favourable for agricultural production (Parren and de Graaj, 1995: 41).

Complimentary Contributor Copy

8

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

COLONIALISM AND THE BIFURCATION OF TRADITIONAL LAND TENURE ARRANGEMENTS Colonialism ushered in transformation and customary and traditional tenures gave way to a modern (statutory) tenure system. This is characterized mainly by the expropriation of community land and forest land and the imposition of state/public ownership (Pougoué and Bachelet, 1982). These customary systems have proven resilient and continue to regulate access to forest land and forest resources. The uneasy cohabitation of land between the modern and the traditional system tenure is faithfully captured by Mahmood Mamdani’s (1996:52) concept of decentralized despotism8. He argues that “the colonial state was ‘bifurcated’ not only spatially but also politically”. On the one hand rural African ‘subjects’ were governed by chiefs and custom and lived on spatially distinct communal lands. On the other hand, ‘citizens’ (whites and other urban dwellers) were governed by modern civil law and owned or rented private property. It was not uncommon for rural colonial officials to assume vacant land as communal property, to be administered and controlled by the chiefs. Colonial administrators therefore granted chiefs enormous control and leverage over land. Land tenure is today characterized by mutually contested rights, what Oyono rightly qualifies as a “deep conflit de langage” (in french)-a conflict of discourse about claims and rights to forest. This conflict of claim and interest, Liz Alden Wily (2012) maintains “directly affects most rural Africans and among whom 75 percent still live on less than US$2 a day.” This situation has created a sense of alienation and deprivation among forest dwellers who are helplessly seeing their resources being carted away by multinational forest exploitation companies with the complicity of the state- leading to a nonchalant and lackadaisical attitude towards the preservation of forest resources. Customary land rights vary across Cameroon. There are two regimes of customary land rights-individual and extended family on the one hand and community rights on the other hand. In the fragmentary societies of Southern and Eastern Cameroon-the seat of the DBR, these rights, are based on the exploitation of forest spaces for agriculture and the demarcation of individual or family estates. Community rights are however exercised on virgin forest as a common property resource. In hierarchical societies which are characterized by a strong, centralized political system as found in northern and western Cameroon, individuals, nuclear families and village communities have bestowed tenure arrangements on traditional authorities for purposes of redistribution, control and the regulation of access to forest. Four characteristics govern these two types of customary forest regimes (Oyono, 2009: 5-6). First , forest land and forest resources are considered sacred. Myths, norms and rules have been put in place to regulate access because in local agrarian logics, land is self-contained. As Anyangwe (1984) rightly maintains, land belongs to itself, and nobody can really own it for himself. As a common property resource, traditional rulers and their Councillors in the Western Grassfields region of Cameroon oversee these forest lands (see Fisiy, 1992) and met 8

Following Mahmood Mamdani (1996) both direct and indirect rule are actually variants of a despotism. While direct rule denied rights to subjects on racial grounds, indirect rule incorporated them into a "customary" mode of rule, with state-appointed Native Authorities defining custom. By tapping authoritarian possibilities in culture, and by giving culture an authoritarian bent, indirect rule (decentralized despotism) set the pace for Africa; the French followed suit by changing from direct to indirect administration, while apartheid emerged relatively later.

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

9

out sanctions to defaulters. Tenure rights are believed to supersede both the individual and the community itself because it has to be preserved for the wellbeing of present and future generations. Secondly, the evolution of customary land tenure systems through history and across economic production systems, resulted from social and political structures that are constantly being reshaped by the latter. These determinant factors account for variations in customary forest tenure systems in rural Cameroon (Oyono, 2009: 5-6). Thirdly, the conflation and lumping together of traditional and modern forest tenure by French and British colonial administrations—under the banner of state ownership—led to new considerations impacting customary tenures, particularly land registration and privatization. As articulated by Le Roy (1982), new forms of property rights arrangements had to ally with the economic and political objectives of colonization, such as capitalist accumulation and the imposition of new types of land and forest governance and administration. In fact, colonialism orchestrated “significant transformations in the meaning of land and private property...” and simultaneously transformed “the existing kin-ordered mode of production into a capitalist economic arrangement, a system that challenges traditional notions of property and land tenure” (Pemunta and Fonmboh, 2010:40). Fourthly, despite the deleterious effect of powerful external forces such as colonialism and the transformation from subsistent to market economies, customary forest tenures have remained resilient and at times adaptive (Oyono, 2009: 5-6). For example, the Batoufan sacred forest in Western Cameroon- an area owned by, and believed to be the abode of the gods to around 100 independent chiefdoms is controlled and guarded through various community-based secret societies. These forests are endowed with rich biodiversity, and each forest possesses different cultural and spiritual values for the communities concerned. Although community institutions regulate access to these forests, community members-particularly sacred healers can enter either to collect key medicines from time to time or through limited annual access, when all community members can enter to harvest a wide range of products. However, immigration to the zone has led to the diversification of cultural norms which tends to whittle the authority of the customary system, as well as conflicting rules between national forest and conservation laws, and customary protection measures and spiritual practices (see Tchouama, 2002). Cameroon’s forest reserves were established by the colonial government in the 1930 and 1940s. Most of these reserves were located north of the country- but have progressively been extended all over the country. Since the 1970s many forest reserves in the humid rain forest zone have been conferred “protected area” status. According to the International Union for the Conservation of Nature (IUCN) (1994) a “protected area” is “an area of land and/ or sea especially dedicated to the protection and maintenance of biological diversity, and of natural and associated cultural resources, and managed through legal or other effective means” usually under several overlapping management regimes. This natural area of land/or sea is designated to (a) protect the ecological integrity of one or more ecosystems for present and future generations, (b) exclude exploitation or occupation inimical to the purposes of designation of the area and (c) provide a foundation for spiritual, scientific, educational, recreational and visitor opportunities, all of which must be environmentally and culturally compatible. The IUCN cautiously accommodates community participation while making conservation the number one priority-but does not rule out forms of sustainable use. This baptism of forested land/and or sea was later followed by the establishment of a number of internationally funded conservation initiatives “set up to assist African governments to

Complimentary Contributor Copy

10

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

manage protected areas” (Malleson, 2000:24) by linking conservation and development. The Integrated Conservation and Development Projects (ICDPs) paradigm is “... an approach that aims to meet social development priorities and conservation goals” (Worah, 2000). The main hallmark of ICDPs is that they seek to address biodiversity conservation objectives through the use of socio-economic investment tools. In Cameroon, these initiatives included the KNP under the auspices of the Worldwide Fund for Nature (WWF) which introduced this concept in the mid-1980s so as to address some of the shortcomings of, and problems associated with "fines and fences” approaches to conservation in PAs. This initiative was viewed as a radical departure from “preservationist approaches to protected area management” (Larson et al. 1998). Despite shifts towards more devolved models known broadly as Community-based Natural Resource Management (CBNRM) which is supposed to “involve institutional reforms and fundamental shifts in power” (Roe et al. 2009:vii), as well as engage forest dwellers as stakeholders in conservation, they have continued to equate conservation with exclusion from PAs. This has resulted in the weakening of broader conservation initiatives aimed at sustainable use of natural resources (Frost et al. 2006). For example, farmers adjacent to reserves still consider wildlife a nuisance. In the case of the Waza National Park, there is a bi-directional interaction. People enter the park illegally to access natural resources and animals go out either pushed by resource degradation inside the park or attracted by local people’s crops or livestock to feed. In the same light, in Paselephants destroy crops, lions kill livestock, birds prey on cereals leading to conflict between local people and park authorities (Bauer, 2003:4). The underlying ideology-the legal prioritization of forest and animals at the detriment of the local population has remained the same in the postcolonial era. As poignantly articulated by Fonjong (2006:670), since independence in 1961, the policies and activities of the GoC in the forest sector seem to replicate the colonial “peopleexclusion-strategy”. This wholesale adoption and replication of colonial era policies of environmental protection whose basis was people-exclusion has failed following Fonjong (2006:672) because: “the context within which the two actors operate are different and therefore, need different approaches. The state, particularly from the late 1980s, operated within the context of growing population, growing urbanization and urban growth, high poverty, harsh economic climate, and emerging [processes of] democratization and globalization. Globalization fosters the growth of international trade. This means for Cameroon an external impetus to deforestation in the form of logging by foreign and multinational companies. Next follows a synoptic overview of the legal and administrative architecture of the land use allocation and land cover types in Cameroon’s national forest estate through June 2011, as well as recent trends in production forests.

THE 1994 FORESTRY LAW: AN INSTRUMENT OF GOVERNMENTALITY AND EXCLUSION The greatest milestone in fortress conservation in Cameroon was the enactment of law no. 94/01 of January1994 (herein henceforth the 1994 Forestry Law) regulating wildlife and

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

11

fisheries and the concomitant decree of application and the 1996 Constitution. This body of forestry and wildlife laws provide for community involvement in natural resource management. Community forestry, public participation in planning and management, and benefit sharing are being promoted to ensure that income from the extraction of forest resources, notably timber, is channelled into development in the production areas. In 2002, the Cameroon National Forest and Environment Sector Programme identified priority conservation areas and forms of forest-land use that could contribute to social development and improved livelihoods (Frost et al. 2006) through poverty alleviation. The 1996 Constitution empowers all citizens to a healthy environment and the right to information and participation in environmental issues (no 96/06 of 18 January, 1996). A framework law defining the orientation for future environmental legislation to operationalize these rights (no 96/12 of 5/8/96) followed. Specific articles promoting conservation include article 26 which states that “the social environment” (meaning culture) of “local communities” must be respected in the creation of PAs. It maintains the right to “normal use” of the area-provided it is not antithetical to the objectives of the local inhabitants. Furthermore, forest dwellers may be entitled to compensation provided their activities have to be discontinued or moved. However, the operationalization of this provision is difficult because many terms remain ambiguous. Ironically, this provision is only applicable to newly created PAs since the law is not retroactive. Theoretically, it grants local people access to new NPs, but there is no precedent as subsistent activities within NPs are still outlawed-even though they continue to take place. Article 29 states that NPs must have a management plan that includes a definition of usufruct rights by local people. In reality, most parks, except Waza and the KNP do not have a management plan (Anonymous, 1997). Section 2.5 elaborates on this management plan and on the inclusion of resource use by local people. In 1995, the enactment decree of the 1994 Forestry Law laid down the procedures for community involvement in commercial hunting and/or wildlife related revenue sharing (see Egbe, 2001) and Mayaka (2003). These however exclude NPs which remain the property of the state. The Forestry Law No. 94-01 of January 20, 1994, and its associated application texts (e.g., Decree No. 95-466-PM of July 1995) laid down the political and strategic framework for forest management in Cameroon. This body of laws define the National Forest Estate (NFE) and subdivide it into two different land use categories—Permanent Forest Estate (PFE) and non-Permanent Forest Estate (nPFE)—each with specific use rights and management regimes. The former (PFE) consists of lands designated to remain as either forest or wildlife habitat. Lands in the PFE are not necessarily forested—many protected areas and hunting zones are located outside forested areas (see Mertens et al. (2012). By law, the PFE must cover at least 30% of the national territory, be representative of the nation’s ecological diversity, and be managed sustainably according to management plans approved by the relevant administrative authority. The non-Permanent Forest Estate (nPFE)—including community forests—consists of forested lands zoned as areas that may be converted into other land uses (e.g., for agriculture). All forests not explicitly held by private entities are under the state’s prerogative. In essence, all forested lands that are not explicitly classified as part of either the PFE or the nPFE fall by default into the nPFE under the category of unclassified state forests (forêts du domaine national). To date, significant areas of the nPFE are still officially classified—these

Complimentary Contributor Copy

12

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

areas consist largely of forested lands under customary use for swidden agriculture or agroforestry purposes (Ibid). In 2011, Cameroon’s NFE represented 37% (17.5 million hectares [ha]) of its total land area. Of this, 94% was allocated within the PFE and the remaining 6% within the nPFE. In terms of land cover, The NFE contains 55% dense forests and 33% mixed forests, and 12% of land cover where forests are not the dominant vegetation. Between 2006 and 2011, the PFE increased by 3% to 16.3 million ha- (35%) of the total national land area, “surpassing the 30% target stipulated by the 1994 Forestry Law”. This increment is attributed largely to “a slight growth in areas classified as either council forests or protected areas”. Fifty-five (55%) of land within the PFE is allocated to production forests (including council forests) whereas 45% to protected areas. The dense forests represents 66% of land cover under the PFE, 11% is mixed forests, whereas 23% occupied by land where forests are not the dominant vegetation. Between 2004 and 2011, the classified land area within the nPFE, witnessed a slight increase of 221%, amounting to 1.1 million ha. The main driver was a significant increase in the area allocated to community forests. This increment was distributed as follows: 90% of the classified lands in the nPFE were allocated to community forests and 10% to sales of standing volume (SSVs). A figure for area under allocation to petits titres was not available for 2011. Alongside unclassified state forests (forêts du domaine national) in the nPFE, as enshrined in the Forestry Law, the nPFE significantly increased to 14.6 million harepresenting 32% of the total national land area. Of this total, about 41% is land covered by dense forests, 59% by mixed forests, and less than 1% by land where forests are not the dominant vegetation (Mertensen et al., 2012:8). Article 24 of the Forestry Law stipulates that PAs or reserves are the private property of the State ‘set aside for conservation, the development and propagation of wildlife, as well as the protection and development of its habitat’. Hunting, except for development purposes as approved by the Minister responsible for wildlife is forbidden. Finally, habitation or other human activities are regulated or forbidden. Contrary to good wisdom,” the law makes no mention of exceptions for indigenous peoples, some of whose members continue, however, to lead a semi-nomadic life in the heart of some of the protected areas, especially (the Pygmies)...” (Nguiffo, 2001: 205). Hunting remains illegal in NPs, but for the theoretical possibility to include it in a management plan as defined in Article 29. A management plan can however, only allow “traditional hunting” of non-threatened animal species (class C animals in Cameroonian forestry law). Any action aimed at tracking, killing or capturing an animal for subsistence consumption using material of plant origin is forbidden. Article 80 specifically frowns against the use of fire, poison, torchlight, and dane guns. In reality, most hunters do use these instruments and increasingly automatic rifles. This implies that hunting cannot be legally prohibited. Article 83 lays down the right to self-defense. This implies that anyone who kills an animal in a bid to protect himself, his livestock or his crops is not to be prosecuted. However, the claim of self-defense must be reported to the appropriate authorities within 72 hours. This provision is unoperational and authorities of the Ministry of Environment and Forestry (MINEF) have consistently maintained sealed lips on this to the local population. The 1994 Forestry law also defines the concept of “buffer zone”. This is defined as the private property of the state in which hunting is forbidden, but other activities such as traditional rituals may be allowed after obtaining permission from MINEF. Buffer zones are

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

13

created by the act of gazzetment of the protected area, and it is not a transition zone between the inside and the outside. Buffer zones are still pending implementation. However, most PAs do have a transition zone-although with specific legal status-variously called “peripheral zone” (Waza National Park), “support zone” (Korup National Park), or “multiple use zone” (Dja Reserve) (Bauer, 2003). Before the enactment of the 1994 forestry law, NPs were wildernesses in theory except for tourism and management was aimed at eliminating-not mitigating any interaction between park and surroundings. As Hans Bauer (2003:23-24) notes: Under favourable economic conditions and repressive leadership, this was fairly effective: few villagers entered NPs. In the 1990s, the new environmental legislation coincided with a severe economic crisis, which was an extra argument for new policies. Park management budgets plummeted, leading to degradation of infrastructure and logistical services. MINEF personnel were considerably reduced, and those remaining saw their salary reduced by 50%. They were not trained in the new styles of management and were no exception to the general rise of corruption, desertion, and neglect of responsibilities in the civil service. The economic crisis made co-management an extra attractive option: in theory devolution of responsibilities implies a reduction of workload, high investment costs could be borne by international donors who were eager to finance fashionable international conservation and development programmes (ICDPs).

As Fonjong (2006) rightly maintains, the economic downturn of the 1980s and the accompanying harsh structural adjustment measures imposed on Cameroon led to a mismatch between economic development and population increase. This “imbalance between population growth and the provision of social amenities” was further accentuated” by the inadequacies of social policies during the period of the structural adjustment program in Cameroon”(UNDP, 1998, cf Fonjong, 2006). In rural areas, the consequence was population instability and natural resource degradation since most of the rural masses depend on agriculture and the exploitation of natural resources- land; water, forest and other related resources were endangered. The forest provides fuel-wood, building materials, wildlife game, parkland, herbs and medicinal plants and is the habitat for other ecosystems and wildlife. Degradation has been further accentuated by poor farming methods, colonization of new agricultural fields, logging for domestic and commercial purposes, and hunting (Fonjong, 2006: 667). The proness of the forest to degradation is compounded by the fact that unlike natural resources such as petroleum or minerals, its exploitation does not require high technological know-how and expertise, neither does human activity in the forest require any specialized skills and huge capital” (Fonjong, 2001). An evaluation of Cameroon’s forestry sector reforms aimed at establishing transparency, equity and ensuring the sustainability of forestry resources shows that most reforms have fallen short of expectations for four reasons. First and foremost, there is lack of genuine commitment by the government and limits on its capacity to implement reforms. Secondly, key stakeholders-particularly foreign logging concessions and parliament are against reforms. Thirdly, partners, such as the World Bank failed to put in place “an implementation strategy that is compatible with the underlying dynamics of political and socioeconomic changes in Cameroon”. Finally, while Cameroon’s body of forestry legislation is well codified in documents, it is poorly implemented (Essama-Nssah and Gockwoski, 2000) due to weak

Complimentary Contributor Copy

14

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

institutional development, inordinate and unbriddled corruption and influence peddling. As noted by Nelson and Gami (2003) weak state structures, highly centralized laws governing the exploitation of forestry resources and lack of accountability over the allocation and regulation of state-sanctioned extraction rights bedeck local communities in the management of their forest resources. The end result has been the rapid and unprecedented depletion of forest and wildlife, even in so-called PAs and NPs. Below we consider two case studies- the KNP and the DBR which represent fortress conservation programs and articulate the problems of NPs and PAs in Cameroon particularly and the Central African region at large.

Case Studies Korup National Park (KNP) Earlier designated as a Native Administration Forest Reserve by Order No. 25 of 14th October 1937, it was in 1986 through Presidential decree No. 86/1283 that the GoC established and extended the KNP to cover a current surface area of 1260 km square (125,600 hectares). Located between latitudes 4° 54’ and 5° 28 North; and longitudes 8° 41 and 9° 160 East (Mbile et al. 2005), the KNP is found between Mundemba and Eyumojock in the South West region of Cameroon. It is contiguous with the Oban National Park in neighbouring Nigeria. It contains the largest number of species of trees in any rainforest in Africa. It has over 400 bird and flora species, 140 fish species and various mammals and primates. More than 620 species of trees and shrubs and at least 480 species of herbs and climbers have so far been recorded. Korup is the single richest lowland site in Africa for birds (more than 400 species), herpetofauna (82 reptiles and 92 amphibians) and butterflies (around 1,000 species). Additionally, there are 130 different species of fish and more than 160 mammals. The KNP is reputed to be richer than any other African forest for which comparable data exists (Richards, 1952). Despite the stress on the ecosystem, the forest has a biomass and production equivalent to other African forests (Newbery et al., 1998). The KNP and forest to the west including the Cross River National Park in Nigeria is reputed as one of the oldest rainforest in Africa (Gartlan, 1984:18). The Korup forest region is characterized by a distinct dry season, from December–February, and rainy season, which peaks between June–October. The mean annual rainfall is in excess of 5000 mm (Zimmermann, 2000, Mbile et al. 2005). The KNP is further characterized by high levels of endemism (Gartlan, 1984:16, Martin, 1991:39). Its high biological diversity is attributed to its survival as a rainforest ‘refugium’ throughout the pleistocene glacial period (Gartlan, 1984:16, Martin, 1991:38). Korup is actually located in the middle of the Guinea Congolian forest refugium (Maley, 1996; Maley and Brenac, 1998). Many fauna and flora species in this park are endangered and some are found nowhere else on earth. The uniqueness of Korup lies in its high degree of endemism amongst its indigenous species (Ruitenbeek, 1990; Thomas, 1986). The area receives a large amount of rainfall and a relatively low amount of sunshine. These factors, combined with poor accessibility, have allowed the natural rainforests to flourish in the area. The park is also known for containing the plant Ancistrocladus korupensis (Ancistrocladaceae), also found in the adjoining Oban

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

15

National Park in Nigeria. The US National Cancer Institute identified this and another plant as capable of providing a cure against HIV that causes AIDS. If the early promise actually translates into real medicinal value, then the plant could provide a source for villagers in the area as well as make a great contribution towards world health.

Source: Draft Management Plan of Korup National Park, 2002. Figure 1. Map of Korup National Park and Support Zone.

Complimentary Contributor Copy

16

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

In 1937 when KNP was created, three legal enclaves existed within its boundaries, to allow the inhabitants of three reserve villages Bera, Esukutan and Bakumba (since abandoned), to have user rights over fishing, hunting and the collection of fruits, food and construction materials (KNP Management Plan, 2002, Mbile et al. 2005). Following the 1961 plebiscite that led to the reunification of then West and East Cameroon, Nigeria became a foreign country and after negotiations between Erat (another reserve village) and the Forestry Department, a legal enclave was established for them in the southern section of the Korup Forest. In October of 1986, Presidential Decree No. 86/1283 established and extended the boundaries of the KNP to include three villages- Ikenge, Bareka Batanga and Ikondo kondo I- technically outside the Korup forest reserve. This decree transformed the livelihood situations of the above-mentioned villages as well as those of the inhabitants of Bera, Esukutan and Erat. With this new legal development, the integrity of the boundaries of this national park became contentious. Simultaneously, there was awareness of a potential threat from 23 villages occurring within three kilometers of its new limits. Despite the presence of five villages with a population of 1500 people inside the KNP and 23 others with an approximate population of 2700 just within three kilometers of its borders, earlier available evidence suggested that very little human disturbance had occurred in the past (Gartlan, 1986). This probably accounts for the park’s initial species richness which is today endangered by human activities. From the date of the raising of the legal status of the Korup forest from a reserve to a national park, the State and its non-governmental partners decided “that, since all the communities inside the park had become illegal residents they should be resettled outright, as it was understood that their close links to the park’s resources were no longer compatible with the conservation and management objectives of the National Park”. Nevertheless, the park managers have always recognized the need to integrate the communities’ perspectives, capacities and wishes in park management. This strategy, within the framework of the overall Park Management program, has come to be referred to as the Community Conservation Strategy (CCS) (Mbile et al. 2005:8). The KNP project is the largest and most famous project undertaken by the World Wide Fund for Nature (WWF), alongside the GoC, Organisation for Development Assistance (ODA), German Technical cooperation (GTZ), EU and WWF-UK. Its purpose is to protect and maintain the park as well as integrating it into regional development plans. The WWF project covers an area of 4,500 square kilometers including the KNP, forest reserves of NtaAli, Rumpi Hills, Ejagham and buffer zones for agriculture, watershed protection and hunting. Today, boundaries have been clearly demarcated; nature trails created, surveillance posts and camp sites built and credible anti-poaching methods have been undertaken by the conservation consortium. The project also includes sub-programs that combine park development and management with environmental education and training. For centuries, the lives of the people in the Korup area have been intertwined with the forest. There are 29 villages in the area, 6 of them inside the park (Malleson, 2000). So the WWF project increases in importance as it tries to provide alternative education and training to the villagers to help reduce poverty and stop them, as well as hunters from all over the country, from indiscriminate poaching of animals. Over the years, this has meant providing facilities for villagers to get training in local technical colleges, employing local people as park guards, providing fertile land outside the park for agriculture, setting up women's cooperatives, providing training in handicrafts, relocating villagers and helping in the development of the villages.

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

17

So far, the results have been quite encouraging for the WWF as many villagers have allegedly “turned in” their weapons, which were earlier used to kill endangered species of animals. Despite efforts on the part of these development agencies to promote the cultivation and marketing of indigenous fruit and medicinal trees, people living in or in the vicinity of (KNP) in Cameroon, are reluctant to engage in tree planting schemes. This reluctance is partly justified by farmers in terms of the limited amount of land available to suit their livelihood needs. In the late 1980s, optimists initially considered KNP a pacesetter in conservation and development. By the late 1990s, it became synonymous with catastrophic failure associated with the integrated conservation and development paradigms due to intense human activities.

The Dja Biosphere Reserve (DBR) The DBR is straddled between the East and South regions of Cameroon, between 2°50 and 3°30 latitude North, and 12°20 and 13°40 longitude East. Covering an area of 5, 260 sq. Km (Betti and Lejoly, 2009), it is classified among the largest PAs of the Guinea-Congolian tropical rain forests. It is located in a transition zone between the forests of southern Nigeria, south-west Cameroon and the forests of the Congo Basin. The humid tropical forest of Cameroon’s Southern and Eastern regions constitute the Congo basin that is sandwiched between Cameroon, Gabon, the Central African Republic and the Republic of Congo. The forest area of the Congo basin covers more than 1,813,000 sq.km. Classified as the second largest tropical ecosystem in the world after the Amazon basin, it holds a capacity of about one-fifth of the world’s tropical forest. It is habitat to over 10,000 plant species, over 1,000 bird species and over 400 mammal species. The area is also home to over 30 million people (Congo Basin Forest Partnership, cf Oke 2009:2). The DBR is secluded by the Dja River, which almost completely surrounds it, and is the habitat to more than 1,500 known plant species, over 107 mammal species, and more than 320 bird species (De Wachter, 1996). At the time of his research, Gartlan (1989) estimated population density at about 1.5 inhabitants/ sq. Km. The diversity of species present in the park, the presence of five threatened species, and lack of disturbance within the park led to its inscription as a UNESCO Heritage Site in 1987. The Bantus and the Baka Pygmies live both inside and outside the reserve. Bantu groups include the Badjoué in the North, the Nzimé in the East, the Mbulu in the West, the Fang-Nzaman in the South, and the Baka Pygmies and the Kako farmers who live in small dispersed settlements, at some distance from the Bantu villages and roads. These people depend entirely on forest resources from the reserve for their livelihood. The Bantus practice “slash and burn” cultivation with a bimodal annual farming cycle, which is entirely dependent on the rainfall pattern (De Wachter, 1996). Both groups practice hunting and gathering, but this activity is more intensively practiced by the Baka Pygmies. A transhumant population of about 25,000 Baka in 42 clans is scattered from the Reserve south into Gabon, northwest Congo and a corner of the Central African Republic. They are mixed among Bantu villages but are unified by the Ubangian language and common customs and beliefs. Except for recent over-hunting of elephants for ivory, the impact of the pygmies on their environment is barely perceptible (Gabon DFC, 2004). By 1999 there were 37 villages with about 3,000 inhabitants living in the Dja Reserve. These poor people depended on 350 medicinal forest plants, rattan cane and raffia (of 4 main genera), wild fruits of which at least 68 species are eaten (ECOFAC Cameroun, 1998), and on hunting.

Complimentary Contributor Copy

18

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

Source: http://rainforests.mongabay.com/congo/. Figure 2. Map of Congo Basin Countries.

There is both subsistence and commercial hunting for deer and monkeys, but intensified elephant hunting by the Baka and by ivory poachers from outside is potentially more disruptive of the traditional order. In 1998, 22,500 people lived along the bounding roads subsisting on fish, while others cultivated cocoa and coffee in a mosaic of plantations within the forest edge (Bedel et al., 1987). Thanks to a cooperation agreement between Cameroon and the European Union (EU) the reserve has been managed since 1992 by the ECOFAC project (Central African Forestry Ecosystems) as one dimension of a regional programme under the auspices of IUCN, at the behest of the European Commission, with the aim of promoting: “the conservation and rational use of forestry ecosystems in Central Africa”. With financing from the European Development Fund, the ECOFAC programme is initiating the establishment of a network of protected areas across Central Africa, with the aim of safeguarding important pockets of biodiversity in the face of exploitation of forest resources in the countries concerned (Mendouga, 1999). In Cameroon, the project seeks to implement initiatives and measures intended to promote the conservation and sustainable use of the reserve’s resources.

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

19

Source: Oke, (2009:2). Figure 3. Map of Dja Reserve.

Inside and around the periphery of the park, theme-based activities are addressing the following issues: post-exploitation development of the forest; instigation of new forms of production; use of available diversity in forest products; rational use of wildlife; research; development of tourism, etc. (Nguiffo, 2001:200). With the management plan still unoperational (UNESCO, 2011), the Zoning Plan put in place by Decree No. 95/678/PM of 18 December 1995, is contrary to traditional Baka practices. It defines, amongst other forest areas, the multiple use zones, where activities by the population are tolerated. Unfortunately, the Miatta Baka, whose customary land rights apply to the Mabé area only enjoy user rights within the Permanent Forest. Within the reserve, they cannot exercise rights to which they are entitled under forest law, since the reserve is outside the multiple use zone, where, the local people can reap the benefits of community forests. Although in the 1995 decree the Forest Zoning Plan is only ‘indicative’, it is unlikely that the boundaries of the reserve will be amended to accommodate Baka claims for two reasons: (1) the boundaries of the Dja Wildlife Reserve were laid down in the founding documentation, prior to the Forest Zoning Plan; and (2) the Dja River, which forms a loop around the reserve provides a natural barrier to the Mabé region for the Baka people of Miatta.

THE PARODY AND FALLACY OF THE GOVERNMENTALITY OF PAS The concept of PAs is rooted in the dual Euro-American strategic models of the 1960s and 1970s (West and Brechin, 1991, Mbile et al. 2005). It is further characterized by two models: 1) “the exclusive management” paradigm that was largely developed in the US consists of separating the interest of local communities from protected areas, with either open

Complimentary Contributor Copy

20

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

anti-participatory attitudes or the outright resettlement of the local communities, 2) “inclusive paradigm” promoted by European states is based on the assumption that guaranteeing the immediate and future “well-being of those who live and work in NPs must always” take precedence (Mbile et al. 2005, Harmon,1991). The former model is ideal for fortress conservation, whereas, the latter is the obvious paradigm for the conservation of PAs that includes human settlement and resource use. Despite ground level reality, there is the predominance of the exclusive model in most of Africa- including Cameroon. Mbile et al. (2005) have bemoaned the fact that the “quasi-exclusive” management regime in the Korup area has led to disaster because of collusion between park guards and illegal poachers from and out of the community. In the Dja Reserve, the situation is the same (see above). Between 1986-2003, despite few estimates of the success rate of this approach, Llewellyn-Smith (1998) estimated that only 20% of the Korup park is adequately protected. Hunting with the use of high caliber weapons (against huge mammals) smuggled from Nigeria remains frequent and extensive. According to the 2002 Management Plan commercial hunting by organized and highly motivated individuals remains high skilled (KNP Management Plan, 2002). With the collusion of village communities or individuals who often receive a fee from ‘outside’ hunters, animals are increasingly at risk of depletion. The situation is compounded by the fact that guards are forbidden from so-called “No- go areas” by the hostility of villagers, which limits the authority of NP guards. The management of both the KNP and the Dja Reserve are increasingly aware that fortress conservation is not only an unsuccessful strategy because of collusion in the community and petty corruption among guards, but like elsewhere in the Less Developed World (De Marconi, 1995), they still regard the local community as the main threat to the park. Unfortunately, the current Management plan of the former suggests increasing the number of guards whereas, the guard’s de jure authority in itself cannot ’protect’ the KNP, as much as they would with community collaboration since community members have more cultural, historical and moral attachment and de facto authority in the use of forest resources in the park. The same situation holds sway for the Dja Reserve. As Mbile et al. (2005:9) rightly maintain: ... the natural resources in the park are vastly more important and central to the villager’s livelihood than to the guard. The villager is also more culturally and historically attached to the resources than most guards (guards drawn from the community are viewed with great suspicion). This combination of closeness to and dependence on the resources, customary proprietorship and indigenous knowledge, gives to the villager’s greater intrinsic moral and de facto authority over the resources of the Park. This de facto use of the park is demonstrated by the extent to which the communities not only collect products and continue to hunt in the park, but farm crops and even permanent crops like oil palm and cocoa.

Different groups of forest users or stakeholders have an isolated perception of the use of the forest which leads to resource use conflict that is ironically not always properly managed. Local dwellers rely heavily on forest products and on agriculture. However, shifting cultivation and slash-and-burn are not only the most widespread agricultural systems in tropical forest, but also constitute the major cause of their destruction (Jepma and Blom, 1992; Cleaver, 1992; Nounamo and Yemefack, 2000). While most governments of

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

21

developing countries depend on the export of tropical wood, logging is inadequately conducted and negatively affects the forest-even in so-called PAs. Although tropical forests are recognized as a real reservoir of biodiversity and other non-consumable benefits, much still needs to be done to preserve them. While conflicting interests are unavoidable, the real task for achieving sustainable development is to harmonize contradictory and unavoidable users' needs (Foahom, 1998). Land for agriculture, non-timber forest products, timber, biodiversity, wildlife, ecotourism and other nonconsumable benefits are important components of sustainable development, but it is also important to ensure the equitable distribution of benefits and address rural poverty. The governmentality of nature through PAs is antithetical and is insensitive to the sociocultural life worlds of diverse forest dwellers such as the Baka around the Dja Reserve and villagers around the Korup area. This tends to generate conflict between conservationists and local communities. First and foremost, the concept of protected area is foreign to the Baka who view ownership of space as belonging to “no one other than the Creator”. For the Baka, the only restrictions to the use of forest products are those related to taboos” (Nguiffo, 2001: 204). Secondly, elephant hunting-the very basis of masculinity in the Baka cultural universe clashes with protected area legislation. Whereas the Forestry law classifies elephants within Class A- one of the most protected species, in Baka culture, elephant hunting is “a statusenhancing activity for the relevant individuals, and every Baka male aspires to kill at least one during his lifetime” (Nguiffo, 2001:204-205). Additionally, the Baka’s traditional system of economic exchanges either with local villagers or with strangers through barter and, increasingly, the sale of forest products that constitute a major source of revenue for the Baka like for villagers around the Korup area have been disrupted since they are contrary to the rules governing user rights since it only permits the taking of forest resources for “personal use”. Additionally, the law on hunting seems to make the Baka’s traditional – and most commonly used – hunting methods illegal, such as the snares made from steel wire, or metaltipped arrows. Simultaneously, hunting with rifles, is now widespread amongst the Baka who receive firearms from their Bantu patrons. Worse of all, the officially decreed hunting period runs from November to February annually. What is pretty clear is that in various cultures, there are festivities that take place at various intervals that require particular plant and animal species. At the same time, ‘the natives’ have their own social calendar based on the activities they undertake at given intervals within the course of the year. The central position here is that forestry conservation laws must be tailored to suit the indigenous calendar. In a sense, ‘the natives’ forest public holidays as well as the national laws put in place turns to reduce the hunting period at the detriment of the former. This is simultaneously a threat to sustainable resource management. The pygmies, as a case in point, have socio-cultural activities to undertake in January, February and March, which include the gathering of mushrooms and hunting. Within the prism of the present legal framework, the vision of the Dja and Korup Reserve conservation projects are fundamentally contradictory to the perceptions of these reserves and their environs. Opposition to development initiatives that could “alleviate poverty” and improve the well-being of the local masses are often framed, but also opposed in terms of environmental protection by environmental NGOs. In their oppositional logic: the forest and the animals come first while the local inhabitants come last since PAs are out of bounds for all communal activities. This prioritization scale begs the question as to who is more important and why are

Complimentary Contributor Copy

22

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

we conserving the environment. In this case, several chasms and contradictions are at stake. There is disregard for indigenous knowledge and participation. Most conservationists tend to neglect the local biodiversity conservation strategies and knowledge systems in favour of a neoliberal business science model. As Johannesen (2006) rightly notes depending on regional economic activity, the expansion of NPs and other PAs may actually reduce both the degree of wildlife conservation as well as local human welfare as the case of the KNP and DBR suggests. Climate change is expected to increase variability, further amplifying these risks. Closely related to this is the clash between local and national interest on the one hand and the global common good. In order of priority, local communities should be the first beneficiaries, but in reality only a few predatory political elites actually benefit from the present arrangement despite talk about community participation in forestry management. Through kickbacks and other incentives, their loyalties are often bought and their mouths sealed. The Cameroon government’s vision of environmental conservation and protection is that of evicting indigenous peoples from their ancestral homelands accusing them of destroying the environment while completely ignoring the fact that they have lived in harmony with and successfully conserved the environment for generations. This nature governmentality strategy in other African countries such as Tanzania and Kenya includes evicting pastoralists from their grazing areas. Despite the pre-and post-colonial alienation of lands, in the name of creating NPs, forest and game reserves, forest dwellers have for some decades now settled in vestigial areas where they have managed to establish villages in accordance with the land laws governing the state. Although their mobility is already constrained by these processes, the government has persistently continued to push them further to the margins, citing investment policies such as the construction of dams for the generation of hydro-electricity, environmental degradation and overstocking, and labelling them as the cause of environmental degradation and conflicts with other land users. This demonstrates that the ideology and interest embedded in the notion of PAs is antithetical to the well-being and human security of forest peoples. Attitudes towards conservation differ among forest peoples. Many communities in wildlife areas are often sidelined and never even consulted and yet they bear the costs of living with wildlife (Kiss, 1990). Consequently, community members have developed a negative attitude towards conservation (Osmondi, 1994; Hill 1998). However, despite the costs of living with wildlife, some communities have maintained a positive attitude towards conservation (Newmark et al., 1993). The rapid depletion of wildlife has been observed in areas where benefits do not accrue to the local community (Norton-Griffiths, 1998). This is because the community engages in other land-use practices that are not only detrimental to wildlife population, but also result in increased conflicts with other resource users leading to the tragedy of the commons. The Bakossi landscape in Southwest Cameroon is a typical example of such an area (Ebua et al. 2011). Additionally, there is the clash between local authority and national authority systems. Generally speaking, the participation of indigenous people in decision-making processes at both the national (macro) and local (micro) levels is limited. There is a fundamental chasm between modern institutions of the nation-state and the traditional indigenous decisionmaking structures, which evolved for different purposes and in most instances are in conflict with each other.

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

23

They are incompatible because the nation-state in Africa opts to impose the notion and practice of the modern state such as modern land tenure on indigenous communities with no respect for traditional systems of governance and authority over land. Indigenous peoples’ perception of the state as an alien authority has led to their reciprocal hostility. As a result, whatever the state stands for is viewed by forest dwellers with suspicion and considered a strategy of domination. This perception is, however, gradually changing as a result of several advocacy initiatives aimed at encouraging states to transform their policies and make them more sensitive to the realities and needs of forest peoples. With the whittling of the powers of the former, the later lays claim to all parcels of land. By parcelling these lands out to multi-nationals in the name of socioeconomic development and poverty alleviation, the state often pits itself against local communities because the concept of community participation or co-management of forests has remained a reality only on paper. With the complicity of corrupt state officials, multinational logging companies never respect the terms of providing social amenities such as hospitals, schools, paved roads and water to the community whose natural resources they are greedily siphoning away. The end result is that of pushing them into deeper poverty and compromising their well-being and those of their children. This state of things has generated conflict between eco-protection and support for the enjoyment of present and future generations. In fact, during the last couple of decades, disguised global investments and modern conservation and climate change policies have become the justification for coordinated violations of human rights and dispossession of forested lands in Cameroon, threatening the future survival of indigenous peoples and their livelihood. It might be stated that in the process of putting an end to the habitation of NPs, the government has consistently used the state machinery and a wide range of national development strategies (Stacey, 2010:24). Against this backdrop, traditional social networks and institutions may break down as communities become less involved in decisions regarding land management, and as decisions made at a higher level replace or obfuscate traditional networks and common pool resources. Some of the most influential changes include the creation of PAs and wildlife/land use policies, the villagization program of the 1970’s, and the expansion of area under cultivation. Land tenure, land use and wildlife regulations of many forms aim to protect natural resources from human use and destruction. In Cameroon, all land and all wildlife belong to the state to be held in trust for the country’s citizens. Population growth also interacts with these changes to impact the per capita land base (Stacey, 2010:4).

INSTITUTIONAL CONSTRAINTS TO INTEGRATED CONSERVATION AND DEVELOPMENT PROJECT (ICDP) This section examines the institutional factors for the dismal failure of the ICDP paradigm in Cameroon and the resulting environmental degradation of NPs and PAs. Institutional weaknesses are the greatest obstacle preventing the implementation of ICDP at NPs and PAs in Cameroon. The administrative institutions in charge of forest and environmental management are highly hierarchical and unadapted to the principles of decentralization.

Complimentary Contributor Copy

24

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

As Samuel and Urie (2011:9) point out these institutions lack cohesion, thereby increasing the number of declarations of intent, which makes it rather difficult to honour commitments taken by the government and therefore limits the coordination of activities. In Cameroon, more than 40% of the PAs created on paper have not been implemented. Cameroon therefore, falls below the objective targeted by the law 94/01 which consists of protecting 30% of the country’s land surface. Matters are further compounded by insufficient financial resources for the implementation of these programmes and projects and therefore strong dependence on external donors and their ideological strings. For instance, the projected budget between 2011 and 2015 of the Ministry of forest and fauna and of that of the environment and protection of nature varies from 17.9 to 20.0 and from 5.9 to 6.7 billion Cfa Francs respectively (Samuel and Urie, 2011). This dependence on external funding has resulted in ill-adapted choices in the Cameroonian context. There is therefore the need to domesticate these choices concerning the community management of forest and natural resources. Furthermore, communities face enormous difficulties in financing and in carrying out the simple management plan and the validation of this plan by the administration. They lack the necessary manpower and expertise. Coupled with this, certain sections of the 1994 forest law contradict the previous laws and regulation but have not yet been repealed (Mvondo and Oyono, 2002). They are inconsistent with the objectives assigned to community forestry. For instance, the 1994 forestry law does not insist on the necessity of protecting certain ecosystems that are endangered by the activities of the forest dwellers. While institutions are important “in influencing the success of governance arrangements”, the success of decentralized forestry management in Cameroon is constrained by the “inappropriate institutional structure at the local level with responsibility to manage community forests” (Brown and Lassoie, 2010: 261). Carolyn Peach Brown and James P. Lassoie further note that: ... prior to the enactment of community forest legislation, included those with historical traditional cultural authority, in the form of clan or lineage heads, as well as the village chief, a legacy of colonial power. Village chiefs or other members of the village council are also selected on the basis of their good moral character (Ibid).

However, due to constraints related to the economic development of the community forest, and due to the meagre benefits that the community activities provide (Lescuyer et al., 2003), the local populations are not really worried by the negative impacts caused by some of their activities on forest biodiversity. They have developed a lukewarm attitude because they were dispossessed of their lands and were rarely consulted. When they were consulted, their grievances were infrequently addressed. On the whole, community involvement and participation has remained a farfetched dream. Local stakeholders’s participation is often limited to attendance at meetings. There is also the exclusion of the voices of women, children and settler groups, even when they are key stakeholders. While these PAs are considered out of bounds for all communal activities, it has led to increased workloads for women and children who have to walk long distances to fetch firewood and water. This exclusion is the result of the trappings of the cultural homogeneity mindset that underpins fortress conservation policy even when most of these

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

25

areas are characterized by heterogeneous livelihood strategies. Consequently, all parks and protected areas are experiencing a high rate of invasion and degradation. Although the invasion rate of the permanent forest by the local populations is relatively low in the forest plantations (approximately 4%), it is quite accentuated in the fauna reserve. In the latter, these rates vary from 35% (in the Dja fauna or wildlife reserve) to 100 % (in the Douala-Edéa wildlife reserve) (NPME, 1996, cf Samuel and Urie, 2011). The participatory approach is further bedecked by differential accumulation of information between stakeholders. This has led to inappropriate public choices that are antithetical to poverty alleviation strategies because finances allocated to impoverished communities within the framework of “integrated” projects are insufficient. This is further compounded by the fact that participants to the different “consultative” meetings are not sufficiently representative of the forest’s traditional owners (Samuel and Urie, 2011). Finally, decisions on the preservation of fauna and flora are handed down in a top-down manner by the central authority. These decisions do not reflect the local needs and aspirations of forest dwellers (see also, Nonga, 2002). This paramilitary approach to conservation through the mechanism of NPs and other PAs does not live up to the objectives of national resource management. This paradigm is stymied by a combination of factors-the impossibility to fence off the PAs, limited logistics (the few reserve staff are poorly remunerated and demotivated and corrupt), illegal exploitation of tree trunks, and frustration resulting from the exclusion of local residents. Against this backdrop, some residents secretly continue with their activities. Lastly, there is the non-consideration of the aspirations of local people and the various sociocultural and symbolic meanings ascribed to the forest by them (see also, Samuel and Urie, 2011:5, Pemunta, 2013). Even more compromising to the alien concept of NPs and PAs is the oversight of the socioeconomic diversity of forest areas. Stated otherwise, there is a gross failure to take note of the diverse social and economic characteristics of forest dwellers and the relationship to their livelihood strategies (Malleson, 2000, 2001). Contrary to bureaucratic beliefs, forest dwellers are often not isolated groups of socially homogenous communities, but rather heterogeneous, cosmopolitan and dynamic people with contrasting livelihood strategies, who are a constitutive element of the regional economy. In the Moloundou District adjoining both the Boumba and Lobéké zones for example, 65% of Baka rely on hunting and gathering as their main livelihood strategy. They spend a significant part of their year in hunting camps far away from the road. Unlike them, those in the Southeast live in close interaction with more settled Bagando who tend to have larger plantations and gardens nearer the main transport axes, and recent migrants, including traders or workers for the various logging concessions in the zone (Ndameau, 2000, WWF, 2002, Nelson, 2002, Tchikangwa, 2002). This diversity of resource use by the surrounding populations suggests that the de facto and de jure situation can be partially reconciled by allowing consumptive use of natural resources at NPs. In Malawi and Uganda for example, the consumptive use of natural resources in NPs has been allowed. In the former country, people are allowed to collect honey inside the Nyika National park insofar as they do not hunt or collect firewood. This has been facilitated by the 1992 Wildlife Conservation and National Park Act which encourages the sustainable consumptive use of national parks (IUCN, 1997). In the later, the consumptive use of natural resources in NPs is formally sanctioned, but regulated. In Lake Mburu National Park, cattle have access to water sources inside the park, fishing is tolerated under certain conditions and people may harvest papyrus, medicinal plants and firewood (Hulme and Infield, 2001). Additionally,

Complimentary Contributor Copy

26

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

emergency access to grazing and water incase of drought or disease outbreak has been arranged (Infield and Namara, 2001). These are pragmatic acts of sustainability because wherever NPs are conceived as pristine wildernesses, there is intruision and misuse leading to degradation and resource depletion (Lowry 1994, Kopas 2008). The current notion of ICDP in the Lobéké National Park is quite exceptional in that contrary to Cameroon law, international conservation NGOs negotiated with and obtained legally, controlled access for local communities to a portion of the park for subsistence purposes. Through community consultations across the zone, conservation agencies such as the WWF, GTZ, and government authorities at MINEF resolved to institute consumptive use. These agencies are determined to ensure that each of the zones in the areas between the parks are attributed to clear stakeholders, and for a better involvement of community representatives in making nominations to the management committees for the conservation areas located adjacent to the park, such as community forests covered by the COVAREF initiative. These committees are working to develop management plans for these zones, which can allow for some sport hunting where the committee can levy a charge. It will also hold potential for the committee to develop a community forestry component, and to control or protect other forms of subsistence hunting and gathering upon which many communities still rely. Currently the system is grossly inequitable, with a clear bias towards minority interests who are not the sole or primary forest stakeholders. For example, only 10% of the local delegates to the management committee of ZICGT 9, a communal forest that adjoins Moloundou Town and the Boumba River, are from the majority Baka community, and are most reliant on the forest to secure their livelihoods (WWF, 2002, Nelson, 2002). In Cameroon NPs are established on lands which local communities had previously been using to secure their livelihood. This act of outright dispossession, broken livelihood and human insecurity has led to intense psychological pains of displacement and loss of identity. For instance, the Baka of Miatta were declared persona non grata from their homeland when they were forced out of the Dja Reserve. They initially lived at Mabé-a village in the heart of the reserve between 1940 and 1950. They were relocated under the aegis of the Cameroon government’s National Sedentarization Policy. They were never consulted and they only came to discover that they no longer have access to their ancestral homeland with its rich fauna and flora. They now suffer from identity crises. However, they still go there regularly to ensure their livelihood-through hunting and the collection of other food products, but more importantly to meet with their co-ethnics from Lomié, Méssaména, Bengbi and Somalomo. The Sedentarization Policy put them at the beck and call of the Bantus who expropriate their labour and have unfettered access to their expertise in traditional medicinal plants-all for a pittance. They lost their traditional authority system and became vassals to the Bantu Chiefs who have become judge and judged in conflict pitting Bantus against Pygmies (Rasek and Schmidt, 1997:18). This legal mirage has led to loss of confidence in both state and Bantu institutions. They are now involved in agriculture-particularly in food producing crops because of their precarious land tenure situation. They are considered as squatters on Bantu lands where they are selling their agricultural labour to their Bantu neighbours for next to nothing (see Nguiffo, 2001: 203-204). Once state laws decree a forest project, the inhabitants of the area have to move. Resettlement usually causes problems as the movement of people to create a dam for instance may give rise to an increase in the incidence of malaria. In such an instance, there is

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

27

the need for an Environmental Impact Assessment report that may predict the incidence of malaria increase in the area and therefore, the extension of a malaria control programme in the region (World Bank, 1999:03). In the process of resettlement, socio-cultural ties/links are /may be severed. Graveyards left behind, as well as other cultural relics and edifices may be a cathedral to the indigenes (see Pemunta, 2013). People are also attached to their physical environment. Particular tree species that thrive only in this region may have ritualistic and medicinal values. This is clearly the case of the Chad-Cameroon pipeline project which wrecked the cultural and socioeconomic livelihoods of the local inhabitants (Ibid). These stringent restrictions on indigenous land use and access apart from eviction of local residents, also involves a moratorium on the consumptive use of natural resources. Worldwide NPs have been established based on western ideologies in non-western systems, generally with the aim of preserving an ecosystem’s local floral and faunal biodiversity in a ‘pristine’ state relatively untouched by human exploitation or occupation (Leader-Williams and Albon 1988), and to provide space for interacting and mutually dependent non-human species (Callicott et al. 1999). But this protectionist ideology does not take into account that many ecosystems now protected (or partially protected) from human use actually evolved with land uses that helped to shape the very ecosystems that the policies were developed to protect. In addition, the loss of access to resources and restrictions on land use that come with the creation of a national park can restrict livelihood options of local inhabitants and other people. As a result, many of the areas outside of PAs in Cameroon have been plagued with poor relations between local people and land management decision-makers since colonial times. Upon the creation of a national park, the displacement of local people creates a direct problem for the evictees as they are re-located, and also a secondary problem for the residents of the area into which they are moved as local population density, and hence pressure on the local natural resource base, is increased. In addition, competition and conflict may ensue if cultural groups are juxtaposed on the landscape. Because PAs are frequently designed to protect key resources such as permanent water, grazing areas, and other resources that local inhabitants historically shared with wildlife, the removal of these areas from local use represents both a loss of area and a reduction in available options. As accessible area is decreased in heterogeneous landscapes with patchy resource arrangements and availability, access to heterogeneity decreases as well (Boone et al. 2000). New options or new locations may be less able to provide resources of sufficient quantity and quality for livestock herd maintenance compared to historical land-use zones. The concentration of pastoralists’ livestock in either smaller areas or in seasonally inappropriate areas taxes the vegetative resilience of these areas, and people’s ability to cope with sporadic drought may be marginalized. In addition, traditional ecological knowledge of the historical land use base is lost (Goldman 2003). The most devastating blow to fortress conservation is the establishment of various capitalist development schemes by the GoC and multilateral donors such as the World Bank. Since colonial times, forest dwellers have continuously been pushed and hounded from their ancestral land to give way to different categories of protected areas, including NPs, Game Reserves and Conservation Areas. PAs have mostly been alienated from forest dwellers. Forest lands have also been alienated for hydro-electricity, mining, foreign and local investments, parastatals and farms, pushing the residents onto marginal areas unsuitable for their livelihood patterns (Stacey, 2010). Two of these projects which infringe on indigenous rights and land without any sustainable environmental safeguards are worthy of examination.

Complimentary Contributor Copy

28

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

The Lom Pangar Dam project-a partnership between the GoC and the World Bank will allow for a major expansion of the Alucam aluminum smelter at the expense of residential consumers and local businesses of electricity. While the Alucam smelter -- owned by mining multinational, Rio Tinto already consumes approximately half of Cameroon’s electricity at low preferential rate compared to residential users, it is further seeking to more than double its production. Lom Pangar will help the government to regulate seasonal flows of the Sanaga River, allowing the construction of more hydro dams downstream, including the Nachtigal Dam set to be built and operated by Alcan for their expansion. It will further increase the vulnerability of Cameroon’s economy to drought, which could be exacerbated by climate change. Cameroon's small businesses and residential customers will remain at risk of blackouts and power shortages. Communities living near Lom Pangar would feel new strains on their already limited resources because of fishermen and others arriving in the area, increased hunting and farming in nearby protected forests, and increases in illness and disease. Furthermore, the dam’s reservoir would flood part of the World Bank sponsored Chad–Cameroon Pipeline and the protected Deng Deng Forest Reserve- home to a community of endangered gorillas. The aluminum industry in Cameroon is clearly being prioritized over the energy needs of the country’s majority population- at great social and environmental risk, and without a participatory planning process for energy development. The Lom Pangar project will flood over 319 square kilometers (nearly 32,000 hectares) of some of the last remaining hardwood forests in Central Africa, including portions of the protected Deng Deng reserve which provides refuge to threatened primates. The project will also endanger biodiversity in the surrounding forests, due to the creation of new access routes into the project area. The dam’s reservoir will submerge sections of the Chad-Cameroon oil pipeline, generating new safety and environmental risks, and will displace hundreds of farmers and herders in one of Cameroon’s poorest regions (Global Village Cameroon, 2006). Cameroon has leased out 73 086 hectares of fertile land in the Ndian and KupeManengouba Divisions of Southwest Cameroon to SG Sustainable Oils Cameroon (SG-SOC further referred to by its parent company-Herakles Farms) for 99 years, for the establishment of a palm oil plantation and palm oil refineries. The annual rent per hectare is scandalously low-50 cents (250 frs CFA) for uncultivated land and 1USD (500 frs CFA) for cultivated land. Moreover, the enterprise can obtain private land title on the land. Beside this case, China and India have obtained 11,000 hectares of land for the cultivation of rice and maize in Cameroon. In the aftermath of the financial crisis, large tracts of land have been bought in Africa, Asia and the Middle East by governments and large firms as a preemptive security measure against future recessions (Zoé, 2012). According to the terms of the land lease contract, SG-SOC will benefit from tax exoneration from the moment that production reaches 10 tons/hectare on at least 3000 hectare for 10 years. The company is further exonerated from paying custom duties and certain social security benefits during the 99 years of the project. It will however pay less than 15% tax on benefits and will carry over its losses from one year to the other indefinitely and without limit so as to reduce imposable benefits (Zoé, 2012). With the advent of the international carbon market, local communities and the Cameroonian government could soon receive compensation for their efforts in carbon sequestration through forest conservation. The SGSOC project destroys the forests in these concessions to plant palm trees, thus reducing the value of the forest and its biodiversity. The contract further empowers the enterprise and not the community and the government to

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

29

benefit from all carbon credits in the zone under concession. Even if the oil palm monoculture project was to be well managed, it will still lead to the loss of biodiversity. The project is situated in an area with high biodiversity, around four protected areas-The KNP, the Rumpi Mountains, the Mount Bakossi and the Bayang-Mbo sanctuary. These protected areas are habitat for scarce and threatened species and also serve as a corridor for the migration of species. It will disrupt the protection and growth of wild fauna. A further threat to NPs and PAs is posed by sustained logging pressure from outside. Most of Cameroon’s NPs and PAs are located in underdeveloped trans-border regions. For example, the KNP is linked to the Cross River National Park in Nigeria. The Lobéké and Boumba National Parks’ proximity to the Central African Republic (CAR) and Northern Congo are associated with intense commercial bush meat and trophy hunting. Threats to the Dja Reserve’s integrity include: commercial hunting, the exploitation for timber and forest clearing for agriculture, and potential mining near the site and the fact that the reserve’s management, planning and surveillance are barely operational. Like in the Korup area, traditional hunting is increasingly being superseded by modern rifles. Logging roads have given rise to poaching for bush meat which is becoming more profitable. Simultaneously, the certification and trafficking of meat (and timber) have grown beyond the control of the Conservation Service in the Dja Reserve and at the KNP and are a major menace to several protected species. Zoo Stuttgart (2000) reported that in just two months, 13 tons of bush meat from the Dja reserve were sold in a village at the northern park border and that at least 300 elephants are killed each year in the area. During 2005-6 some 1,500 pieces of bush meat were seized (UNESCO, 2006). In 2006 alone, Eco-guards reportedly seized 89 elephant tusks, 517 duikers, 338 blue duikers, 143 bush pigs and 188 monkeys of illicitly harvested species (World Heritage Report on the Dja, 2006, cf, Oke, 2009:24). A Survey conducted by Zoo Stuttgart showed that personnel of the French logger Pallisco rented guns to hunters in exchange for part of their catch which was then ferried to markets on timber lorries. Apart from the service being hugely understaffed and underfunded, law enforcement is weak and officials are intimidated if they want to control transport vehicles suspected of carrying bush meat. Some are actually entangled in the bush meat trade and issue “bogus hunting permits” with no legal basis (Zoo Stuttgart, 2000). Stronger enforcement is needed to register the origin of timber and bush meat, and to monitor the exit roads for illicit export. There are also increasing markets for wildlife and ivory, the elephants falling to Baka hunters as a traditional trophy. Improvement of antipoaching laws and their effective enforcement are still required, as is awareness of the damage that hunting is doing to the Reserve’s potential for ecotourism. There is a need to fund more effective surveillance and to legalize and integrate the Village Watch Committees into anti-poaching programmes and help them with the work (UNESCO, 2011). Furthermore, the GoC-licensed logging of the buffer zone forests, managed by Unités Forestières d’Aménagement, is isolating the Dja Reserve from other forests and is not popular with the locals: dues are paid to them to make up for lost use of their trees, but roadside treecutting, hunting and looting of forest resources are facilitated by loggers’ tracks and villagers are sometimes defrauded over the dues. Communes inside the core zone are paid no compensation and need an alternative source of income such as ecotourism (UNESCO, 2006). An EU grant of US 50,000000 meant to “renovate roads” according to the investigations of the Rainforest Foundation instead led to increased timber exploitation because of improved access, and the discovery of 9 logging concessions (cf Zoo Stuttgard, 2000).

Complimentary Contributor Copy

30

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

Forest inventories have also been undertaken in the Dja area like at the KNP contrary to existing legislation because it contains useful species, therefore, the threat of commercial scale logging remains. Agricultural encroachment which is slight at present will grow as the population grows wherever the border demarcation remains unclear. Traditional slash and burn cultivation for cocoa, coffee and village subsistence plots within the reserve- particularly on the northern and western borders pose further threats. Like at the KNP, a larger threat looming over the Dja reserve is industrial scale plantations, now developing on the western periphery. Better coordination with these surrounding interests is needed as well as within the Conservation Service itself (UNESCO, 2006, 2011). Mining activities by the authorized GEOVIC Cameroon Company are even more devastating. Miners have been developing deposits of cobalt, nickel and manganese 40 km east of the border. A legal, environmental and social impact study of the concession is still required, so as to gauge the potential for heavy pollution from settling ponds and of the potential impacts of 700 workers and 2.000 incomers. The pollution could affect the health of the local populations, especially the Baka pygmies. The Trans-African highway, though on the other side of the river, runs close to the southern boundary of the Reserve. The Reserve was recently explored for oil and gas, and the existence of calcareous bodies on the south-east border of the Reserve could lead to open-cast mining for cement production (UNESCO, 2011). Several safari companies operate in the area, and several natural resource goodies including live parrots, ivory and other illegally obtained forest resources are regularly obtained in or smuggled through the zone. Zoo Stuttgard estimates that within 60km radius of Lomié, at least 6 gorillas, elephants and chimpanzees are shot every month and there are 27 hunting camps up to 50 km into the reserve. These produce an estimated 10 tons of bush meat per week (Zoo Stuttgard, 2000). Added to this is the tragedy of the 880km Chad-Cameroon oil pipeline undertaken by the World Bank, Exxonmobile, Shell and Elf Aquitaine which is close to the border of the reserve. The World Bank’s justification was poverty alleviation in both Chad and Cameroon. It has instead led to pauperization. Some tracks lead through pristine forests and settlements of forest dwellers, animal tracts (e.g. of elephants) were severed, approximately 2,500 workers cut wood for building their huts and fed on wildlife in the surrounding forest (Zoo, Stuttgard, 2000, see also Pemunta, 2013). Worse still, these trans-border regions are underdeveloped and characterized by state-centric, top-down development with rudimentary economic growth.

POLICY IMPLICATIONS While the creation and implementation of NPs as artifacts and processes illustrate the nature and extent of global governmentality upon these regions, their people and their natural resources, it highlights the chasm between law as a governmental rationale which often fails to recognize traditional usage rights and creates conflicts between PAs and local communities. As Messe Venant rightly notes, Cameroon created PAs in response to a specific needthat of enabling the government to conform to international agreements. For forest and wildlife managers at the local level, substantive negotiations over enabling traditional

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

31

activities inside PAs often seem impossible, because government representatives ‘are trying to apply the law as is their duty’ (Venant, 2009:9) and as if these laws are cast in stone. However, laws present only a partial or fragmentary picture of who controls access into the forest. In their application, they are often contested, circumvented and selectively applied, interpreted and re-interpreted in their making and implementation (Krueger, 1974, Bates 1981, Bhagwati, 1982). There is a whole network of extra-legal, social, cultural and politicoeconomic relations in which law is embedded. These shape the effects and meaning of any laws or legal changes- whether of electoral codes or of environmental regulations (Ribot, 1995, Comaroff and Roberts, 1981). Common experience shows that political-administrative laws that systematically disable accountable local representation often frustrate participatory rural development efforts. The concept of community participation in Cameroon remains problematic. Participation must involve the exercise of political power and locally accountable representation. Unfortunately, all over Cameroon, power elites have highjacked this representation for their own selfish interests. At times, they are the ones who acquire the status of community forestry for their people. Representative bodies ought to have real powers of decision-making over valuable resources. Even when Chiefs and councilors take it upon themselves to represent in a meaningful way the local community, their representativeness and accountability remains problematic. Decentralization, “the devolution of central state assets and powers to local/private decision-making bodies: representative local government, local administrative branches of the central government, non-state, [voluntary organizations] (NGOs, co-operations, associations etc.) or private individuals and corporations” (Jesse and Oyono, 2005:45) can be one form of community participation. At times, they actually represent only their stomachs and family members at the detriment of the local community. They tacitly connive with the administration to dispossess their own people. Transparency and accountability at all levels of governance are essential prerequisites for developing better natural resources management. Clearly defined but also dynamic sets of cultural rules, values and belief systems form and inform the traditional institutions of governance and decision–making, which influence and regulate the day-to-day relations within households and within the entire community in indigenous communities. These indigenous communities operate on dual system of governance and decision-making. On the one hand, there is the traditional system of governance based on social values, norms and institutions, which are mainly defined, in varying degrees, from community to community, by men. Then, on the other, we have modern administrative structures of governance and leadership that are based on the modern laws, human rights and democratic values used by the government to regulate day-to-day decision-making. In most cases, the two tend to clash or have divergent approaches (Porokwa, 2009). In Cameroon like in most of Africa, customary and community-based systems for managing land rights are incompatible with state laws. This is the case even when most people hold land under customary law, and their rights to do so are almost never recognized under existing national laws. Among the Bulu of Cameroon land tenure is based on the “droit de hache” principle. Original settlers/the first to exploit a parcel of land or their kin assert proprietary rights over that parcel of land, and allocate it according to customary rules. This hierarchical framework restricts land access to kin, and strangers must negotiate access. Land right is inheritable and land can be portioned out among users- “rights to cultivate versus rights to harvest tree products”.

Complimentary Contributor Copy

32

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

Additionally, “the scope and strength of rights may be tempered by the land typology; rights to cropped lands are more stringent than those to high forest”. Contrary to the Bulu, the mainly hunters and gatherer Bagyeli “Pygmy” communities living in close association with them have a different land tenure system. Their customary tenure concepts revolve around rights to lands where they cohabit with agrarian and trading communities. They however assert clan rights to certain portions of the forest and high forest, but grant usage rights to outsiders on conditions that yields are shared and that they comply with conditions of use (Nelson and Gami 2002:3). Cameroon’s 1994 Forest Law defines the right to land through the mise en valeur principle. This restricted definition of land as “being under exploitation” –as being used for agriculture contradicts longstanding and sometimes contradictory arrangements of socially sanctioned methods of land acquisition, use and transfer making ethnic groups such as the Pygmies “virtual land owners”. While the mise en valeur principle is embedded in national law, forested lands fall under a government protected and regulated category which may be allocated to commercial and conservation interests- often to outsiders. Contradictions between modern and local tenure rules of this nature has given a lee-way to the potential destruction of century-long customary land tenure arrangements as observed in many communities in Cameroon. By stipulating that all land that is not held under private land title belongs to the State, Cameroon’s land tenure law excludes the Pygmies and other forest peoples with different livelihood strategies. This law is at variance with customary rules, which define the principle of land acquisition, usage and transfer between all the different forest peoples. The overlap between modern and customary law is detrimental to the Pygmies, who are transformed into ‘virtual owners’ of land through their use of it., but whose rights are not recognized and guaranteed under modern law. In fact, Cameroonian law and jurisprudence favour modern law over customary rules (Pemunta 2013:362-3).

Disjunctures between the State “and local rules of this nature allow for the potential destruction of customary land tenure arrangements that have been operating for generations, a fate which has befallen many communities in Cameroon” (Nelson and Gami, 2003:3). This has resulted in growing land insecurity among indigenous communities such as pastoralists and hunter-gatherers, who already face serious social marginalization and pressures from external agricultural, logging, mining and conservation interests. This disjuncture between state laws and customary, often communal rules over lands is contributing to significant longterm negative impacts on the livelihood security of indigenous Cameroonian communities, especially with the absence of the reforming processes required to bolster recognition for customary law in formal legal texts. There is the urgent need “to reform laws and their application through the “re-institutionalization” of customary arrangements into codified law” (Nelson, 2004). This will reduce tenure insecurity for indigenous communities, and ensure equity and justice regarding the allocation and management of lands over which these communities have strong claims. The resilience of customary land tenure shows the irreconcilable gap between the dictates of national law and ground level reality which is best illustrated by the overlapping state and community tenure over public land (Wily, 2012) although state laws often take precedence. Customary rights to land should become statutory rights of customary ownership. The new land laws of several African countries- Mozambique

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

33

(1997), Uganda (1998), Tanzania (1999) and Southern Sudan (2009) recognize and provide for integrated plural legalism. This will put an end to the subordination and suppression of customary tenure as a legal means of land ownership. In other words, the change of the law on land tenure will be a tacit recognition of the fact that customary rights to land have the force of modern real property, whether registered or not. Additionally, women should be empowered to own and dispose of land, not only de facto beneficiaries with usurfruct rights over land by virtue of their membership of particular kinship group. Fortress conservation does not reflect ground level reality. The KNP and the Dja reserve are not unique in having people with divergent and “several livelihood clusters” resident within them. Park residency is common in other parts of the world, such as Bolivia (De Marconi, 1995), the United Kingdom (Harmon, 1991), and Canada (Kopas, 2008, Lowry, 1994, 1999). Unlike elsewhere, the main difference with the KNP and Dja Reserve could perhaps be the non-recognition of both the legitimacy and the contribution that such inhabitants can make towards the management of the Park. Conservation should not be simply about protection- creating a terra nullius (a no man’s land)-rather, it should be about the reallocation of resources and the restructuring of social institutions. Conservationists and policy makers often gloss over differences in needs, cultures, and customs- whereas, multiple “livelihood clusters” (functional units of people who make their living in similar ways) occur at the village level simply because forest dwellers live in the same ecological niche, they assume that only groups that are geographically close together share the same interests. Whereas, in reality “several livelihood clusters” exists within each community- different groups have different interests, livelihood strategies, demographics, networks, and interactions with ecosystems”. In the Gariep Basin in South Africa, for example, there is considerable variation in people’s interactions with and connectedness to ecosystems. Although wage labor, remittances, and migration have replaced agriculture as the main sources of income, most people still maintain a link to ecosystems by owning some livestock, cultivating crops, and harvesting fuel wood for heating and cooking (SAfMA Livelihoods) (Cundill and Shultz, ND). This calls for an integrated development paradigm that meets the needs of diverse actors and stakeholders. The creation of a Conservation Development Authority with the mandate to initiate private sector partnership and also encourage community development and participation in the effective management and control of PAs is a necessity. In fact, Ndumu Fidelis Oke makes the point for autonomous management, increased public participation in decisionmaking and enforcement and the creation of a special fund for collected charges from logging companies so that the funds be used to increase conservation efforts as a mitigating step (Oke, 2009). The mandate of this body must incorporate active community participation in decision-making and planning for the sustainable use of ecosystem services and development of ethno-tourism, if trends in rural emigration and depopulation are to be halted and the NP is to be protected in line with sustainability principles. This is one way of meaningfully engaging local communities so as to ensure the goals of conservation. As a governmental regime, the establishment and gazzetment of a protected area alone is insufficient to conserve biodiversity (Bauer, 2003, Burkey, 1995). They are however, an important strategy if their location is well chosen (Myers et al. 2000).

Complimentary Contributor Copy

34

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

Community participation can be achieved through genuine collaboration and mutual respect between conservationists and local communities in the management of PAs and not by the former imposing on the latter. Genuine Community involvement of resource users calls for ‘linking with local residents’ (Mbile et al. 2005)- by avoiding information asymmetry, providing information on policies and environmental problems, supporting decision-making institutions and instruments, and supporting small income generation activities. Both the conservation authority and the locals should sit on the management board as equals as obtains in Australia. Community-Based Natural Resource Management should go beyond the interpretation held by “government authorities, donor agencies, and NGOs as benefit-sharing or outreach between national parks and adjacent communities”. Under such circumstances, ‘‘communities are not empowered as authorized local resource managers but are involved principally as passive recipients of benefits controlled elsewhere. This form of outreach and benefit-sharing is also a characteristic of some protected area management in East African countries” (Roe et al. 2009:vii). Successful community forest management depends on the existence of effective institutions (Pagdee et al. 2006), yet the institutions put in place in Cameroon during decentralization are ineffective in achieving significant socioeconomic or environmental goals-partly because of the usurpation of power from traditionally recognized power holdersthe imposition of inappropriate institutions that has marginalized those who have recognized customary powers and legitimacies over forests, while privileging those to whom decentralization reforms have given power (Oyono 2005, Brown and Lassoie, 2010:266). Indigenous knowledge, long neglected and perceived as an obstacle to development is presently seen as central in discussions of sustainable resource use and balanced development (Brokensha, 1980, Rao and Ramana, 2007). Indigenous knowledge is unique and is at the heart of decision-making in agriculture, health care, education, natural resource management among others. It is valuable for culture of origin and to planners striving to improve conditions in local communities (Warren, 1991, 1992). This important natural resource can facilitate the development process in a cost-effective, participatory and sustainable way. Indigenous technical knowledge about conservation needs to be documented because it is both ecologically and socially significant (Posey, 1985) for its effective use in sustainable development. Roa and Ramana (2007:131) have demonstrated how the Chenchus of Nallamalai forests in India have used traditional knowledge to sustainably conserve their environment for generations. Among others, they do not kill pregnant animals, thereby conserving nature in their own interest. Similarly, they frequently abandon a portion of the tubers or roots in the ground when they dig for food thereby facilitating the regeneration of the roots and tubers. Sustainable forest management (SFM) is pure wishful thinking without an intersectorial and a holistic development approach. Threats to NPs and PAs with rich biodiversity include: immigration resulting in high demand for land and forest products, over-exploitation of wildlife (commercial poaching and trans-border poaching), unsustainable and illegal logging, weak and opaque governance of forest royalties and excruciating poverty. In other words, it is possible only within the context of an overall strategy to promote sustainable development in all sectors of the economy and to alleviate rural poverty. This is in line with Chapter 11 of Agenda 21, of the 1992 Earth Summit. SFM is one of the key frameworks necessary to reach sustainable use of forest biological diversity.

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

35

In this respect, Forest Principles stress that "Forestry issues and opportunities should be examined in a holistic and balanced manner within the overall context of environment and development, taking into consideration the multiple functions and uses of forests, including traditional uses, and the likely economic and social stress when these uses are constrained or restricted, as well as the potential for development that sustainable management can offer" (Sands et al. 1994:875). It was certainly with this in mind that while drawing up its National Forestry Action Plan (NFAP), Cameroon shifted emphasis from the tree to the entire forest and harmonized users' needs (sometimes contradictory), but also attempted to make it coherent and complementary to the national development plans of other sectors. Despite this, implementation has been stymied by several factors-particularly the centralized politicoadministrative system-despite moves towards decentralization as well as inept and unbridled corruption. Industrial logging - the monopoly of foreign consortiums spearheaded by the French is also taking its toll on Cameroon’s forest estate. In 1998 and 1999, excluding other forest related assets, French companies alone held more than 50% of forest concessions, (The Courier, 2002). Most commercial exploitation occurs outside the legal framework provided for by the 1994 forest law that ought to regulate the abusive rate of deforestation and promote reforestation. The illegal exploitation and other malpractices account for the majority of the logging. Oftentimes, public officials thwart the course of justice and defaulters of forest regulations to go scot-free. Global Forest Watch, for example, observed that ‘… one out of every five violation reports have been dropped in the wake of intervention by influential officials …’ (The Courier, 2002).

CONCLUSION The creation of PAs is an act of global governmentality. This study argues that by denying people benefits and access from natural resources, they develop negative attitudes and engage in anti-conservation activities thereby threatening the future of wildlife especially large mammals. In order to promote peaceful co-existence of residents and wildlife, efforts must be made to reverse anti-conservation practices and activities and community members should realize the benefits of wildlife. This is achievable through constructive dialogue and consultations-one that takes into consideration the local point of view and local conservation practices. Ebua et al. (2011) after a year of participant observation in the Bakossi area of the Southwest region of Cameroon have shown that conservation can be successful within and around PAs. In line with them, we maintain that to change the perception and attitudes of indigenous people around these areas, environmental education through sensitization should be encouraged. Negative attitudes and perceptions can be reversed through well designed and carefully implemented conservation programmes. The various shortcomings of fortress conservation articulated in this chapter reflect the failure of law at the international level or their ineffective translation and domestication at the local level. For instance, the UN framework convention on climate change sees trees as carbon-sink or reservoirs and not as the basis for the conservation of forestry biodiversity. As such many happy negotiators agree to the planting of millions of hectares of tree monocultures as a “solution” to the green house effect while most biodiversity negotiators would rightly see this as the worst possible approach.

Complimentary Contributor Copy

36

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

Secondly, it shows the predominance of global and national policies geared to benefit transnational corporation’s interest for whom biodiversity conservation is either an obstacle to profit-making (e.g. pulp and timber sectors) or something to be robbed and patented (e.g. biotechnology and pharmaceutical sectors). Local populations- the true conservators stand on their way and become the victims of the take over and eventual destruction of nature by transnational corporations (TNCs). For the state and forestry consortiums, the forest is a source of private accumulation of economic interest regardless of the interest of forest dwelling communities. For the indigenous population, on the other hand, the forest is the life wire of their survival. Thirdly, the predominance of the market-oriented approach, where biodiversity is worthless unless it has a price tag attached to it and unless such price tag is equal or superior to other possible profitable activities (Blaikie and Simo, 1998) equally impacts negatively on the environment and on PAs. Fourthly, it shows the separation of production from biodiversity conservation. PAs are very useful. They serve no purpose if made up of mere islands within a sea of biodiversity destruction. Most national policies seem to be going in that direction, converting entire landscapes into seas of agricultural and tree monocultures, while thinking that biodiversity conservation can be achieved by preserving some representative areas such as NPs and PAs.

REFERENCES Amman, K., Pierce, J. (1995). Slaughter of the Apes: How the Tropical Timber Industry is Devouring Africa’s Great Apes. World Society for Protection of Animals. London. Anyangwe, C. (1984). “Land Tenure and Interests in Land in Cameroonian Indigenous Law.” Cameroon Law Review, 21, 29-40. Anonymous. (1997). Plan directeur d’aménagement du parc national de waza. MINEF, Yaounde. Arsène, N. (2012). Huile de Palme: Une société américaine au cœur d'un scandale. L'Indépendant. Available at http://www.cameroon-info.net/stories/0,37075,@,huile-depalme-une-societe-americaine-au-c-ur-d-un-scandale.html. Last accessed 09/13/2012. Aubréville, A. (1949). Climats, Forêts et Désertification de L’Afrique Tropicale. Paris: Société d’Éditions Géographiques, Maritimies et Coloniales. Bates, R. (1981). Markets and States in Tropical Africa. CA: University of California Press. Bauer, H. (2000). Lion Conservation in West and Central Africa. Integrating social and natural science for wildlife conflict resolution around Waza National Park, Cameroon. Ph. D Thesis. Institute for Environmental Sciences. Leiden, The Netherlands, 160 pp. Bedel, J., Bousquet, B., Gourlet, S. (1987). Réserve Biosphère du Dja. Report to the Government of Cameroun and UNESCO/MAB by l’École Nationale du Génie Rural des Eaux et des Forêts (Montpellier), 96 pp. Betti, J. L., Lejoly, J. (2009). Contribution to the knowledge of medicinal plants of the Dja Biosphere Reserve, Cameroon: Plants used for treating jaundice. Journal of Medicinal Plants Research., 3,1056-1065.

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

37

Bikié, H., Collomb, J., Djomo, L., Minnemeyer, S., Ngoufo, R., Nguiffo, S., Ndjodo, T., Ndoye, O., Clark, L., Selig, E., Bryant, D. (2000). A Global Forest Watch Cameroon Report. Washington DC: World Resources Institute. Bhagwati, J. N. (1982). “Directly Unproductive, Profit-Seeking (DUP) Activities” Journal of Polit. Economy,. 90, 988-2002. Blaikie, P., Simo, J. M. (1998). Cameroon’s Environmental Accords: signed, sealed but undelivered. Yaounde: National Printing Press. Boone, R., Galvin, K., Smith, N., Lynn, S. (2000). "Generalizing El Nino effects upon Maasai livestock using hierarchical clusters of vegetation patterns." Photogrammetric Engineering and Remote Sensing., 66, 737-744. Brokensha, D. (1980). Indigenous Knowledge Systems and Development. University Press of America, Lanham, US. Brown, H. C. P., Lassoie, J. P. (2010). Institutional choice and local legitimacy in community-based forest management: lessons from Cameroon. Environ. Conservation., 37, 261–269. doi:10.1017/S0376892910000603. Available http://environment.upei.ca/ files/environment/brown__lassoie_env_conservation.pdf. Last accessed 14/09/2012. Burkey, T. V. (1995). Faunal collapse in East African game reserves revisited. Biol. Conservation, 71, 107-110. Callicott, J., Crowder, L., Mumford, K. (1999). "Current normative concepts in conservation." Conserv. Biology., 13, 22-35. Callister, D. J. (1999). Corrupt and illegal activities in the forestry sector: current understandings and implications for World Bank Forest Policy. Paper prepared for the World Bank. http://wbln0018.worldbank.org/essd/forestpol-e.nsf/MainView. Last accessed 09/13/2012. Cleaver, K. (1992). Deforestation in Western and Central African forest: the agriculture and demographic causes and some solutions. Proceedings of the Conference on Conservation of West and Central African Rain Forest, Abidjan, 5-9 November 1990. World Bank Environment Paper No 1. The World Bank, Washington DC, US. Colchester, M. (1994). Slave and enclave: the political ecology of equatorial Africa. World Rainforest Movement, Penang, Malaysia. Comaroff, J. L., Roberts, S. (1981). Rules and Processes: The cultural logic context. Chicago: University of Chicago Press. Convention on Biological diversity. (1997). Report. Republic of Cameroon. United Nations Environment Programme. http://www.cbd.int/doc/world/cm/cm-nr-01-en.pdf. Last accessed 09/13/2012. Cundill, G., Schultz, L. (2007). Communities, Ecosystems, and Livelihoods. Mario Giampietro, Thomas Wilbanks, Xu Jianchu. (Eds), pp. 261-277. Available at http://www. maweb.org/documents/document.349.aspx.pdf. Last accessed 09/13/2012. Davenport, T., Usongo, L. (1997). Justification and recommendations for the gazettement of a “protected area” in Lobeke Forest South East Cameroon. Report prepared for the Ministry of Environment and Forests, Cameroon. Lobeke Project, World Wildlife Fund, Gland, Switzerland, and Global Environment Facility, World Bank, Washington, D.C., US. De Marconi, M., (1995). In: Amend, S., Amend, T. (Eds.), Inhabitants of protected areas in Bolivia: National Parks Without people? The South American Experience. IUCN, Quito.

Complimentary Contributor Copy

38

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

Djeumo, A. (2001). The development of Community Forests in Cameroon: Origins, Current situation and Constraints. Rural Development Forestry Network, 25, 1-16. ECOFAC Cameroun. (1998). Spécial Réserve du Dja. Petit tour d’horizon de la biodiversité du Dja. Canopée No.12. De Wachter, P. (1996). Economie et impact de l'agriculture itinérante Badjoué (Sud Cameroun). Civilisations, 44 (1-2), 62-93. Ebua, V., Buh, A., Evaristus,T. F., Ngudem, S. (2011). Attitudes and perceptions as threats to wildlife conservation in the Bakossi area, South West Cameroon. International Journal of Biodiversity and Conservation., 3, 631-636,. Available online at http://www.academic journals.org/IJBC. Last accessed 09/13/2012. Egbe, S. E. (2001). The law, communities and wildlife management in Cameroon. RDFN Paper 25e(i),1-12. URL http:www.odi.org.uk/publications/rdfn/256rdfn-25e-i.pdf. Last accessed 09/13/2012. Elliot, L. (1998). The Global Politics of the Environment. First edition, London: Macmillan Press. Essama-Nssah, G., James, J. (2000). Cameroon-Forest sector development in a difficult political economy. Environmental Evaluation Country Case study Series. The World Bank: Washington, DC. Ewah, J. O. (2010). Expanding mandate and corporate responsibility in the management of National parks and Protected areas in Nigeria. Journal of Agriculture and Environment for International Development., 104(1-2), 25-38. Fairhead, J., Melissa, L. (1998). Reframing Deforestation: Global Analyses and Local Realities: Studies in West Africa. London and New York: Routledge. Fabricius, C., Folke, C., Cundill, G., Schultz., L. (2007). Powerless spectators, coping actors, and adaptive co-managers: a synthesis of the role of communities in ecosystem management. Ecology and Society 12., 29. [online] URL: http://www.ecologyandsociety. org/vol12/iss1/art29/ FAO. (2010). Situation des forêts dans le monde. Rome, FAO. FAO. (2000). Food and Agriculture Organization of the United Nations, statistical databases, Online at:http://apps.fao.org/lim500/nphwrap.pl?Forestry.PrimaryandDomain=SUA andserv let=1. FAO. (2002). Forêts du Cameroun. Retrieved from http://www.fao.org/forestry/fo/country. FAO. (1997). State of the World’s Forests. FAO, Rome. Fisiy, C. F. (1992). Power and Privilege in the Administration of Law: Land Law Reforms and Social Differentiation in Cameroon. Research Reports 1992/48, Leiden: African Studies Center. Foahom, B. (1998). The scientific approach to sustainable forest management, with particular attention to land-use planning and social aspects. In: Tropenbos Seminar Proceedings. "Research in Tropical Rain Forests: Challenges for the Future", 115-123. Tropenbos Foundation, Wageningen, The Netherlands. Foahom, B. (1996). Analyse des implications des Conventions et Accords internationaux relatifs au secteur forestier au Cameroun. MINEF, Cellule d'Etudes et de la Planification. Yaoundé, Cameroun, 28pp. Foahom, B., Jonkers, W. B. J. (1992). A Programme for Tropenbos research in Cameroon. Final Report. The Tropenbos Foundation, Wageningen, The Netherlands.

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

39

Fonjong, L. N., 2001, Fostering women’s participation in development through non governmental efforts in Cameroon. Geographical Journal, 167(3), 223–234. Fonjong, L. N. (2006). Managing deforestation in anglophone Cameroon: are NGOs pacesetters? International Journal of Environmental Studies, 63., 663–679. Frost, P., Campbell, B., Medina, G., Usongo, L. (2006). Landscape-scale approaches for integrated natural resource management in tropical forest landscapes. Ecology and Society., 11: 30. [online] URL: http://www.ecologyandsociety.org/vol11/iss2/art30/. Last accessed 09/13/2012. Gabon DFC (Direction de la Faune et de la Chasse) (2004). Proposition d’Inscription de Biens sur la Liste de Patrimoine Mondial. Écosystème et Paysage Culturel du Massif de Minkébé, Republique du Gabon. Ministry of Waters, Forests and Environment, Libreville. Gartlan, S. (1989). La conservation des écosystèmes forestiers du Cameroun. Gland, Suisse et Cambridge, Royaume-Uni; UICN. Gartlan, J. S., (1986). The Biological Importance of the Korup Forest, In: Gartlan, J. S., Macleod, H. (Eds.), Proceedings of the Workshop On Korup National Park, Mundemba, Ndian Division, SW Province, Republic of Cameroon. WWF/IUCN Project 3206. Gartlan, J. S. (1984). Korup regional management plan: conservation and development in the Ndian Division of Cameroon (Draft). Wisconsin: Wisconsin Regional Primate Research Centre. Global Village Cameroon. (2006). In Whose Interest? The Lom Pangar Dam and Energy Sector Development in Cameroon. A report based on field work prepared by Global Village Cameroon, Bank Information Center and International Rivers Network. Available http://www.internationalrivers.org/files/attached-files/whoseinterest.pdf.Last accessed 09/13/2012. Global Forest Watch. (n.d). http://www.globalforestwatch.org/english/#brief Goldman, M. (2003). "Partitioned nature, privileged knowledge: community-based conservation in Tanzania." Development and Change., 34, 833-862. Harmon, D., (1991). In: West, P. C., Brechin, S. R. (Eds.), National park residency in developed countries: the example of Great Britain. In: Resident Peoples and National Parks: Social Dilemmas in International Conservation., 33-39, University of Arizona Press, Tucson, US Huges, R., Flintan, F. (2001). Integrating Conservation and Development Experience: A Review and Bibliography of the ICDP Literature. Biodiversity and Livelihood Issues No 3. International Institute for Environment and Development. Hobart, M. (1993). An Anthropological Critique of Development: The Growth of Ignorance. Rutledge, London. Hutter, C. (2000). (Ed). An Overview of Logging in Cameroon. Global Forest Watch. World Resource Institute, Washington, DC. Hulme, D., Infield, M. (2001). Community Conservation, reciprocity and park-people relationships In: African Wildlife and Livelihoods, D. Hulme and Murphree, M. (Eds), 106-130. Oxford: James Currey Ltd. Infield, M., Namara, A. (2001). Community attitudes and behaviour towards conservation: an assessment of a community conservation programme around Lake Mburu National Park, Uganda. Oryx., 35, 48-60.

Complimentary Contributor Copy

40

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

Infield, M. (1988). Hunting, Trapping, Fishing in Villages within and on the Periphery of the Korup National Park. WWF report. Washington, DC. IUCN. (1994). Guidelines for Protected Areas Management categories. IUCN Gland URL: http://wcpa.iucn.org/pubs/pdfs/IUCNCategories.pdf. Last accessed 09/13/2012. Jepma, C. J., Blom, M. (1992). Global trends in tropical forest degradation: the Indonesian case. Policy simulation with the help of IDIOM. IDE Discussion Paper 9104, Schiermonnikoog, The Netherlands. Kimbi, A. N. (2012). Attainment of MDGs through tourism in Central African sub-region: Implications for local economic development in Cameroon. Revista Turismo y Patrimonio Cultural., 10, 3-16. Kiss, A. (eds.). (1990). Living with wildlife, Draft report of World Bank Environment Division, The World Bank, and Washington, DC. Kopas, P. S. (2008). Taking the Air and change in Canada’s National Park. Vancouver: University of British Columbia Press, Korup National Park Management Plan. (2002). Ministry of Environment and Forests. Cameroon 2002. Krueger, A. O. (1974) “The Political Economy of Rent-Seeking Society”. American Economic Review., 64, 291-303. Laporte, N., Goetz, S. J, Justice, C. O., and Heinecke, M. (1998). “A new land cover map of central Africa derived from multi-resolution, multi-temporal AVHRR data.” International Journal of Remote Sensing., 18, 3537-3550. Larson, P. S., Freudenberger, M., Wyckoff-Baird, B. (1998). WWF Integrated Conservation and Development Projects: Ten Lessons from the Field 1985-1996. Washington DC. Leader-Williams, N., Albon, S. D. (1988). "Allocation of resources for conservation." Nature, 336, 533-535. Le Roy, E. (1982). “Les objectifs de la colonisation française ou belge.” In: Encyclopédie Juridique de l’Afrique, Kouassigan, Guy, A. (Ed), 85-95. (Tome Cinquième: Droit des Biens). Abidjan and Dakar: Les Nouvelles Editions Africaines. Llewellyn-Smith. (1998). Final Report to WWF-Cameroon. Consultancy Contract Ref. CM 0008.01: Korup Project Park Management Technical Adviser. Lescuyer, G., Nlom, J. H. (2003). Rapport final. Une analyse économique au service des acteurs locaux. Projet COAIT – IRM et CIRAD Forêt. Logo, P. B., Bikie, E. (2003). “Women and Law in Cameroon: Questioning Women’s Land Status and Claims for Change” In: Women and land in Africa: culture, religion and realizing women’s rights, 35-55, Wanyeki, L. M. (ed). London: Zed Books. Lowry, W. R. (1999). “Providing Intergenerational Goods: Implementation of National Park System Plans in Canada and the United States” Policy Studies Journal,. 27, 328–346. Lowry, W. R. (1994). “Paved with Political Intentions: The Impact of Structure on the National Park Services of Canada and the United States”. Policy Studies Journal., 22, 4458. Lynne, S. (2010). The Pastoral to Agro-Pastoral Transition in Tanzania: Human Adaptation in an Ecosystem Context. Available: http://www.economics-of-cc-in- tanzania.org/images/ Stacy_Lynn_Pastoralism_TZ_Draft_2010_08-09_draft_2_v2.pdf. Last accessed 09/13/ 2012.

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

41

Malleson, R. (2000). Forest Livelihoods in Southwest Province, Cameroon: An evaluation of the Korup Experience. Ph.D Thesis. Department of Anthropology. University College London. Malleson, R. (2001). Opportunities and Constraints for ‘Community-Based’ Forest Management: findings from the Korup Forest, Southwest Province, Cameroon. Rural Development Forestry Network paper, 25g, 11-19. Maley, J., Brenac, P. (1998) Vegetation dynamics, Palaeoenvironments and Climatic changes in the Forests of West Cameroon during the last 28,000 years BP. Rev. Palaeobotany and Palynology, 99,157-187. Maley, J. (1996). The African rain forest – main characteristics of changes in vegetation and climate from the Upper Cretaceous to the Quaternary. Proceedings of the Royal Society of Edinburgh. Section B. Biological Sciences, 104, 31-73. doi:10.1017/S0269727000 006114. Mamdani, M. (1996). Citizen and Subject:Contemporary Africa and the Legacy of Late Colonialism. Princeton: Princeton University Press. Mayers, J., Bass, S. (2004). Policy that Works for Forest and People: Real Prospects for Governance and Livelihoods. International Institute for Environment and Development (IIED): London. Mayaux, P., Archard, F., Malingreau, J.-P. (1998). “Global tropical forest area measurements derived from coarse resolution satellite imagery: a comparison with other approaches.” Environmental Conservation., 25:37-52. Mbile, P., Vabib, M., Mebokac, M., Okonc, D. J., Arrey-Mbod, J. F., Nkonghoe, F., Ebongc, E. (2005). Linking management and livelihood in environmental conservation: case of Korup National Park, Cameroon. Journal of Environmental Management. 76, 1-13. Mbile, P., Degrande, A., Okon, D. (2003). Integrating Participatory Resource Mapping and Geographic Information Systems in Forest Conservation and Natural Resources Management in Cameroon: A Methodological Guide. EJISDC http://www.ejisdc.org., 14, 2, 1–11. Mendouga, L., (1998). Comment les populations forestie`res ge`rentelles les 1000 frs par metre cube? FTPP Bulletin 15/16, 16 – 22. Mertens, B., Neba, G. S., Steil, M., Tessa, B. (2012). Interactive Forest Atlas of Cameroon Version 3.0. Overview Report. World Resource Institute. MINFOF. (Ministry of Forestry and Wildlife). (2011). Protected Areas in Cameroon. Yaounde: Government of the Republic of Cameroon. Muchaal, P. K., Ngandjui, G. (1995). Sectuer Ouest ce la Réserve de Faune du Dja: Evaluation de l’Impact de la Chasse Villageoise sur les Populations Animales et Propositions d’Aménagement en vue d’une Exploitation Rationelle. ECOFAC/MEF. Youandé. Mvondo, A. S., Oyono, P. R. (2002). Community based forest management in the mount Cameroon region: case study of implementation hurdles and risks for sustainability. Cifor, Cameroon, 12p. Myers, N., Mittermeier, C. G., da Fonseca, G. A. B., Kent, J. (2000). Biodiversity hotspots for conservation priorities. Nature, 403, 853-858. Neba, N. E., (2009). Ecological Planning and Ecotourism Development in Kimbi Game Reserve, Cameroon, J. Hum. Ecol., 27, 105-113.

Complimentary Contributor Copy

42

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

Nelson, J., Gami, N. (2003). Enhancing equity in the relationship between protected areas and indigenous and local communities in Central Africa, in the context of global change. CEESP-WCPA-IUCN Theme on Indigenous and Local Communities, Equity and Protected Area. Final Report. Forest Peoples Programme Moreton-in-Marsh, UK. Nelson, J. (2004). A survey of indigenous land tenure in sub-Saharan Africa. In: Land Reform, land settlement and cooperatives. Edited by P. Groppo., 64-79. FAO. Available ftp://ftp.fao.org/docrep/fao/007/y5407t/y5407t00.pdf. Last accessed 09/13/2012. Newmark, W. D., Manyanza, D. N., Gamassa Deo-gratias, M. (1994). The conflict wildlife and local people living adjacent to protected areas in Tanzania: human density as a predictor: Conservation. Biology., 8, 249-255. Newmark, W. D., Leonard, N. L., Sariko, H. I., Gamassa Deo-gratias, M. (1993). Conservation attitudes of local people living adjacent to five protected areas in Tanzania: Biolog. Conservation., 63, 177-183. Nguiffo, S. (2001). One forest and two dreams: the constraints imposed on the Baka in Miatta by the Dja Wildlife Reserve. Case study 6 – Cameroon – Dja Wildlife Reserve. Available http://www.forestpeoples.org/sites/fpp/files/publication/2010/08/cameroondjaeng.pdf. Last accessed 09/13/2012. Nounamo, L., Yemefack, M. (2000). Shifting cultivation in the evergreen forest of Southern Cameroon: farming systems and soil degradation. Final Report. Tropenbos Report 00-2. Tropenbos Cameroon Programme, Kribi, Cameroon. 54 pp. Norton-Griffiths, M. (1998). The economics of wildlife conservation policy in Kenya. In: Conservation of Biological Resources, Ed. E. J. Milner Gulla and R. Mace, 279-293. Oxford: Blackwells. NgoNonga, F. (2002). Gestion soutenable de la forêt tropicale et développement intégré au Cameroun. Thèse de Doctorat d’État, Université de Yaoundé II-Soa, Yaoundé, 385 pp. Oke, N. F. (2009). Deforestation Impacts on biodiversity conservation in the Dja Biosphere Reserve, Cameroon. Masters Thesis in Ecotechnology and Sustainable Development. Mid Sweden University, Sweden. 39 pp. Osmondi, P. (1994). Wildlife-Human conflicts in Kenya: Integrating wildlife conservation with human needs in Masai Mara region; PhD Thesis, McGill University, Montreal, 338 pp. Oyono, P. R. (2005). Profiling local-level outcomes of environmental decentralizations: the case of Cameroon’s forests in the Congo Basin. Journal of Environment and Development 14., 1–21. Oyono, P. R. (2009). New Niches of Community Rights to Forests in Cameroon: Tenure Reform, Decentralisation Category or Something else? International Journal of Social Forestry (IJSF), 2, 1-23. Pagdee, A., Kim, Y.-S., Daugherty, P. J. (2006). What makes community forest management successful: a meta-study from community forests throughout the world. Society and Natural Resources., 19., 33–52. Parren, M. P. E., De Graag, N. R. (1995). The quest for natural forest management in Ghana, Cote D’Ivoire and Liberia. Waginengin (Topenbos)., 13,1-1999. Pemunta, N. V. (2013). The governance of nature as development and the erasure of the Pygmies of Cameroon. Geo Journal Volume 78, (2): 353-371. doi:10.1007/s10708-0119441-7 Key: citeulike: 10114913.

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

43

Pemunta, N. V., Fonmboh, N. M. (2010). "Experiencing Neoliberalism from below: The Bakweri Confrontation of the State of Cameroon over the Privatization of the Cameroon Development Corporation". Journal of Human Security., 6., 38-53. Porokwa, E. (2009). Pushing, Hounding and Bullying. Half a decade of resentment and acrimony towards indigenous peoples in Tanzania. Special Issue: Pastoralism. Indigenous Affairs, 3- 4/109,. 2229., http://www.iwgia.org/graphics/SynkroLibrary/Documents/ publications/Downloadpublications/IndigenousAffairs/IA%2034_2009/IA%203_09%20p astoralism.pdf. Last accessed 09/13/2012. Posey, D. A., William Balee (Eds.). (1985). Resource Management in Amazonia: Indigenous and Folk Strategies. Advances in Economic Botany, 7, Plenum Press. New York. Rasek, A., Schmidt, J. (1997). ‘Analyse comparative des systèmes de production agricole Baka et Bantu de la région de Djoum', CED, Yaoundé. Rao, V. L. N., Ramana, G. V. (2007). Indigenous Knowledge, Conservation and Management of Natural Resources among Primitive Tribal Groups of Andhra Pradesh. Anthropologist Special, 3, 129-134. Anthropology Today: Trends, Scope and Applications. Veena Bhasin and M. K. Bhasin, Guest Editors. Ribot, J., Oyono, P. (2006) “The Politics of Decentralization.” In: Towards a New Map of Africa, Ben Wisner, Camilla Toulmin, Rutendo, Chitiga (eds.), 205–28., London: Earthscan. Ribot, J., NY. (1999). Decentralization, Participation and Accountability in Sahelian Forestry: Legal Instrument of Political Administrative Control, Africa, 61:1. Ribot, J., NY. (1995). From Exclusion to Participation: Turning Senegal’s Forestry Policy Around? World Development, 23, 1587-99. Ruitenbeek, J. H., (1990). Economic Analysis of Tropical Forest Conservation Initiatives; Examples from West Africa. WWF Report. Roe, D., Nelson, F., Sandbrook, C. (eds). (2009). Community management of natural resources in Africa. Impacts, experiences and future direction. Issues No. 18. London: International Institute for Environment and Development. Samuel, N. P., Urie, J. E. (2011). Approaches of Forest Sustainable Management and Economic arbitration in Africa. Journal of Sustainable Development in Africa, 13, 1-18. Sands, P., Richard, T., Weiss, M. (1994). Documents in international environmental law. Manchester: Manchester University Press. Sanjayan, M. A., Shen, S., Jansen, M. (1997). Experiences with Integrated-Conservation Development Projects in Asia. World Bank Technical Paper No. 38. Washington DC. Schumacher, E. F. (1973). Small is Beautiful: A Study of Economics as if people Mattered, Abacus edition, London, Sphere Books. (First Published 1973, London, Blond and Briggs). Scott, J. C. (1998). Seeing Like a State. London: Yale University Press. Seignobos, C., Iyebi-Mandjeu, O. (Eds). (2000). Atlas de la province Extreme-Nord Cameroun. Ministerè de la Recherche Scientifique et Technique, Yaounde, and Institut de Recherche pour le développement, Paris. Shakleton, A., Campbell, B., Wollenberg, E., Edmunds, D. (2002). Devolution and community-based natural resource management. Creating space for local people to participate and Benefit. Natural Resources perspectives, ODI. Sillitoe, P. (1998). The Development of Indigenous Knowledge: A New Applied Anthropology. Current Anthropology, 39, 223-252.

Complimentary Contributor Copy

44

Ngambouk Vitalis Pemunta and Ogem Pascal Mbu-Arrey

Stromayer, K., Eboko, A. (1991). A Biological Survey of Southeastern Cameroon. European Union, Wildlife Conservation Society. New York. Tchouamo, I. R. (2002). Constructed Management of Conflicts in Locally Protected Areas: Case of the Batoufam Sacred Forest (West Province of Cameroon). Mimeo. University of Dschang, Cameroon. The Courier, 2002, No. 193 July–August. The magazine of ACP-EU Development Cooperation, Brussels. Thomas, L., Middleton, J. (2003). Guidelines for Management Planning of Protected Areas. IUCN, Gland, Switzerland and Cambridge, UK. UNDP. (1995), 21 Human Development Report (New York: UNDP). UNDP. (2003). Global environment Outlook N° 3. UNESCO. (2011). Dja Faunal Reserve Cameroon. World Heritage Sites, protected areas and World heritage. Available http://www.unep- wcmc.org/medialibrary/2011/06/29/ 53bce643/Dja%20Faunal%20Reserve.pdf. Last accessed 09/13/2012. UNESCO/IUCN. (2006). Rapport de Mission Suivi de l’Etat de la Conservation de la Réserve de Faune de Dja en République du Cameroun, Site de Patrimoine Mondial. Paris and Gland, Switzerland. United Nations Organisation. (1987). Report of the World Commission on Environment and Development: Our Common Future (Brundtland Report). New York: United Nations. Worah, S. (2000). International History of ICDPs. In: UNDP. Proceedings of Integrated Conservation and Development Projects. Lessons Learned Workshop, June 12-13, 2000, Hanö: UNDP, World Bank/WWF. Venant, M. (2009). Securing Indigenous Peoples’ Rights in Conservation: Reviewing and promoting progress in Cameroon. FPP series on Forest Peoples and Protected Areas. Translated from the French and edited by John Nelson, Africa Policy Advisor, FPP. Association Okani and Forest Peoples Programme. Warren, D. M. (1995). The Cultural Dimensions of Development, Indigenous Knowledge Systems. Intermediate Technology Publication, London. Warren, D. M. (1992). Indigenous knowledge, bio-diversity conservation and development. Keynote address. International Conference on Conservation of Biodiversity in Africa, Kenya. Warren, D. M. (1991). Using indigenous knowledge in agricultural development. World Bank Discussion Paper No. 127. The World Bank. Washington D.C. West, P. C., Brechin, S. R. (Eds.) (1991). Resident Peoples and National Parks. University of Arizona Press, Tucson, US. WRM, 1997, WRM’s bulletin No. 4, September. WWF. (2002). “Community protected natural areas in the state of Oaxaca, Mexico”. WWF Forests for Life and WWF Mexico Programme. White, L., Abernathy, K. (1997). A Guide to the Vegetation of the Lopé Reserve. Libreville: ECOFAC. Wily, A. L. (2012). Customary Land Tenure in the Modern World Rights to Resources in Crisis: Reviewing the Fate of Customary Tenure in Africa - Brief #1 of 5.Available at http://www.rightsandresources.org/documents/files/doc_4699.pdf. Last accessed 09/13/ 2012. Worah, S. (2002). “The challenge of community-based protected area management”. Parks, 12, 80–90.

Complimentary Contributor Copy

The Tragedy of the Governmentality of Nature

45

Wolman, M. G., Fournier, F. G. A. (eds.) (1987). Land transformation in Agriculture. John Wiley and Sons, Chichester. Zimmermann, L., (2000). A comparative study of growth and mortality of trees in caesalp dominated lowland African rainforest at Korup, Cameroon. Zoé, J. S. (2012). Ndian/Koupé-Manengouba: Le Cameroun brade des terres à 250 F CFA l’hectare. L'Actu. Available at http://www.cameroon-info.net/stories/0,37056,@,ndiankoupe-manengouba-le-cameroun-brade-des-terres-a-250-f-cfa-l-hectare.html. Last accessed 09/13/2012. Zoo Stuttgard. (2000). Sense and Nonsense of external funds. Example Dja Wildlife Reserve, Cameroon. Available at http://www.bushmeat-campaign.net/engsite/pdf/dja.pdf. Last accessed 09/13/2012.

Complimentary Contributor Copy

Complimentary Contributor Copy

In: National Parks Editor: Johnson B. Smith

ISBN: 978-1-62618-934-8 © 2013 Nova Science Publishers, Inc.

Chapter 2

LINKING PROTECTED AREA CONSERVATION WITH POVERTY ALLEVIATION IN UGANDA: INTEGRATED CONSERVATION AND DEVELOPMENT AT BWINDI IMPENETRABLE NATIONAL PARK J. Baker1, R. Bitariho2, A. Gordon-Maclean3, P. Kasoma4, D. Roe5, D. Sheil6, M. Twinamatsiko7, G. Tumushabe8, M. van Heist9 and M. Weiland10 1

Balfour Beatty, King Hill, West Malling, Kent, UK Institute for Tropical Forest Conservation (ITFC), Kabale, Uganda 3, 5 International Institute of Environment and Development, London, UK 4 Jane Goodall Institute (JGI) Uganda, Entebbe, Uganda 8 Advocates Coalition for Development and Environment (ACODE), Kamwokya, Kampala, Uganda 10 People and Conservation, Kampala, Uganda 2, 6, 7, 9

1. INTRODUCTION Priorities for Managing National Parks Gaining the support of local communities for conservation and resolving local conflict issues are priorities for managers of national parks. Conflict can be defined as the expression of divergent interests between resource-poor households neighbouring a national park and the national and international actors concerned with conservation of biological diversity (Blomley 2003). Conflict can arise when access to natural resources is prohibited or from humanwildlife conflict. In addition to gaining the support of local communities, managers must also protect endangered wildlife and ecosystems from activities that threaten the conservation status, particularly unauthorized resource use. However achieving the balance of improving relations with local communities while enforcing conservation law can be a significant challenge, particularly at national parks surrounded by high populations of rural communities whose livelihoods depend on the natural resource base.

Complimentary Contributor Copy

48

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Poverty Alleviation The 2011-2020 Strategic Plan for the Convention of Biological Diversity (CBD) heralded a new era of national park management when it set an agenda for biodiversity conservation to contribute towards poverty eradication. The 10th Conference of Parties encouraged parties to ‘support initiatives on the role of protected areas in poverty alleviation’ (Decision X31) and, in doing so, identified national parks as important for CBD signatories to deliver conservation-poverty alleviation goals. National park managers must therefore seek to reduce the poverty of local communities using interventions that achieve conservation goals. However linking conservation with poverty alleviation is more than effective national park management, but requires that issues of governance, human rights, equity and power are addressed at the highest levels. This requires governments to align conservation and development policies whereby conservation policies take account of social justice and development policies incorporate environmental needs, and establish a framework that provides conservation and development practitioners with one strategic direction on the governance of natural resources.

Governance Natural resource governance can be defined as ‘the interactions among structures, processes and traditions that determine how power and responsibilities are exercised, how decisions are taken, and how citizens or other stakeholders have their say in the management of natural resources - including biodiversity conservation’ (IUCN Resolution RESWCC3). Commonly recognised elements of good governance include: transparency; access to information; access to justice (and a way of resolving conflict and disputes when they occur); involvement in decision making (indicated by participation, legitimacy and the ‘voice’ that people have); fairness; coherence; performance; subsidiary; respect for human rights; accountability; and rule of law, which should be fair, transparent and consistently enforced (Borrini-Feyerabend et al. 2004). Good governance of natural resources is the foundation for linking national park conservation with poverty alleviation. Achieving good governance within a national park context includes the effective participation of informed local communities in natural resource management, negotiated agreements between communities and authorities on natural resource use, fair compensation for the costs of conservation and equitable benefit sharing that addresses the needs of the poor and marginalised.

Integrated Conservation and Development Integrated Conservation and Development (ICD), where conservation is achieved by addressing local development priorities (Wells et al. 1992), is a tool for national park managers to link conservation with poverty alleviation. While national park conservation is the goal, the ICD approach achieves this through economic development and by providing local people with alternative income sources that do not threaten natural resources (Brandon and Wells, 1992).

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

49

First Generation ICD When first introduced, ICD was considered a radical new approach that held great promise for overcoming major challenges to national park conservation, particularly for developing countries. Integrated conservation and development programmes (ICDP) attracted considerable funding and were rapidly implemented across the world. Early ICDPs were projects that integrated natural resource management with grass-root economics (Larson et al. 1997). In practice social services including schools, health clinics and roads were provided to improve local attitudes towards conservation and, by doing so, reduce threats to the national park. For example, conservation efforts for the Beza Mahafaly Special Reserve in Madagascar incorporated various development activities that included constructing a school and developing a community health programme (Larson et al. 1997). However ICDPs became large, multi-institutional efforts that relied on external expertise. Concerns soon arose over the long-term funding requirements (Kremen et al. 1998) and, as the interventions bore no relation to conservation, that ICDPs were too focused on rural development (Wells et al. 1992). The programmes were widely considered as large, complex experiments that alienated communities from resource management (Kremen et al. 1998) and failed to link conservation and development (Wells et al. 1992).

Second Generation ICD In response to criticism, the ICD approach was refined. Based on the premise that local populations will commit to conservation when their socio-economic well-being is assured (Kremen et al. 1998), the aims were to provide communities with sustainable economic alternatives to unsustainable harvesting and land use practices (Wells and Brandon, 1993; Alpert, 1995) and resolve conflict between national park authorities and local communities. Collaborative management agreements for local resource use were promoted as strategies to address conflict through the sharing of benefits from conservation and decision-making powers among stakeholders (Wells and Brandon, 1993), and as a mechanism to involve local people in natural resource management (Borrini-Feyerabend, 1996). Agreements for local resource use were commonly implemented through a system of buffer zones. The first were adjacent to national parks (Mackinnon et al., 1986) and then became harvest zones inside national parks (Wells and Brandon, 1993). Several projects in tropical forests implemented harvest zones for the collection of minor forest products including wild plant resources, honey and bamboo (Boot and Gullison, 1995). This provided rural communities with vital basic needs such as building materials, fuel, food and medicines, and the opportunity to continue cultural traditions (Cunningham, 1996). These ICD projects varied in size and budget from a small marine park in Haiti with a budget of several thousand dollars, to national level support for ICD in Namibia, which involved $10 million over 10 years (Larson et al. 1997). However, the criticism continued from ecological, socio-economic and governance perspectives.

Complimentary Contributor Copy

50

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Ecological Concerns Ecological concerns included the harvesting of non-timber and minor forest products from national parks. In theory harvesting practices were the least harmful extractive use of forests (Jacob, 1988). However, in practice there were instances where resources were overexploited. For example the destruction of medicinal plants and dye resources by ring-barking and uprooting in Africa (Cunningham, 1987; 1990), the depletion of copal and rattan resources in the Philippines (Conelly, 1985) and the over-exploitation of two species of palm fruits in the Peruvian Amazon (Vasquez and Gentry, 1989; Peters, 1990). While the likelihood of over-exploitation depends on supply, the part of the plant harvested and growth form, an increase in demand for resources by local harvesters was the common cause of overexploitation (Cunningham, 1996). Therefore although collaborative management agreements proved successful in involving local communities in resource management and gaining local support for conservation, the agreements must be based on regulations on the harvesting with the number of harvesters balanced against the conservation value of the species that is harvested (Cunningham, 1996; Scott, 1998).

Socio-Economic Concerns The use of economic benefits as a conservation tool is a common feature of ICD and many strategies have been promoted as providing economic benefits while securing conservation. The sharing of tourism revenue is common at sites where charismatic species attract large numbers of tourists. This non-consumptive means of generating local income is to build national park support by transferring economic benefits to local communities as a means to offset local costs of conservation (Wunder, 2000; Walpole and Goodwin, 2001; Walpole and Leader-Williams, 2002). Revenue sharing can improve local attitudes towards conservation (Archabald and Naughton-Treves, 2001). However, success has been mixed and several reviews have identified that more must be done to link economic benefits directly to national park conservation (e.g. Wells and Brandon, 1993). Distribution issues are a common barrier, particularly the decision of who receives the revenue and how it is disbursed equally. One solution is to share revenue with communities who most immediately affect, and are affected by, the national park (Wells, Brandon and Hannah, 1992; Western and Wright, 1994; Ross and Wall, 1999). However, those who have the greatest impact on conservation are not necessarily the same as those suffering the greatest cost, and the uneven distribution of costs and benefits impedes efforts to ensure that revenue sharing funds achieve conservationpoverty linkages by reaching the poor and marginalised (Barrett and Arcese, 1995; Archabald and Naughton-Treves, 2001).

Governance Concerns Participation is fundamental to ICD yet many projects failed to devolve natural resource management to local communities (Ghimire, 1994). An internal WWF review found that many ICDPs had not incorporated the interests of key stakeholders and that participation was particularly difficult in forest projects where local resource use is intensive (Larson et al.

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

51

1997). Forest projects managed through a centralised body have also been criticised for failing to meet conservation goals and the needs of local people (Fisher, 1995). While participation of local communities in national park management has increased, the type of participation is rarely defined (Box 1), yet critical for conservation managers to evaluate progress towards achieving good governance as an output of ICD. Box 1. Types of participation (adapted from Adnan et al. 1992) Type of participation Passive participation Participation in information giving Participation by consultation Participation for material incentives Functional participation Interactive participation

Self-mobilisation

Characteristics People are told what is going to happen or has already happened. These are unilateral announcements that do not listen to people’s responses. People answer questions posed by extractive researchers and they are not able to influence proceedings, with research findings not being shared with them. People are consulted, but external professionals largely define both problems and solutions. Decision-making is not shared, and professionals are under no obligation to take on board people’s views. People provide resources, for example labour, in return for food, cash or other material incentives. People form groups to meet predetermined objectives related to the project. Such involvement tends to be during later project cycle stages after major decisions have been made. People participate in joint analysis, which leads to action plans and the formation of new local institutions or the strengthening of existing ones. These groups take control over local decisions so people have a stake in maintaining emerging structures or practices. People take initiatives independent of external institutions. They develop contacts with external institutions for the resources and technical advice they need, but retain control over how resources are used.

The ICD Debate Therefore the ICD approach aims to meet development priorities and conservation goals, with the use of socio-economic tools as a conservation strategy. It has proven to improve community-park relations although its effectiveness in linking conservation and development has been questioned (Wells, Brandon and Hannah, 1992; Malleson, 2002). This could be because efforts to reconcile conservation and development are most likely to achieve a best compromise and only problems are documented (Hughes and Flintan, 2001), or that the slow and complex process of changing the way people manage resources and earn their livelihood means that ICD develop and improve gradually (Larson et al. 1997; Abbot et al. 2001; Browder, 2002). The debate as to whether ICD can conserve national parks through poverty alleviation is limited by the lack of multi-disciplinary monitoring of ecological and socio-economic impacts (Larson and Svendsen, 1995). Many ICD evaluations highlight the need for empirical evidence on drivers of conflict and resource use to better target ICD interventions (Blomley et al. 2010). Improving ICD therefore requires a greater understanding of the social, economic

Complimentary Contributor Copy

52

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

and wellbeing profiles of individuals who instigate conflict and harvest resources, and their motivations for doing so. For ICD to successfully achieve conservation through poverty alleviation, there is also a need to account for governance, as monitoring and evaluation efforts tend to focus on conservation and development outcomes rather than indicators of good governance. This limits our understanding of how best to achieve good governance when implementing ICD and, consequently, ICD success. Questions including whether local communities were effectively engaged with the decision-making process, felt a sense of ownership of natural resource management, received fair and equitable compensation for the costs of conservation need to be examined with scientific rigour to fully understand the governance issues that underpin ICD success.

2. LINKING CONSERVATION AND POVERTY ALLEVIATION AT BWINDI IMPENETRABLE NATIONAL PARK ICD at Bwindi

Figure 1. Location of Bwindi Impenetrable Forest.

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

53

The ICD approach was adopted at one of Uganda’s most prestigious national parks – Bwindi Impenetrable National park (hereafter referred to as Bwindi) in south-west Uganda (Figure 1). Bwindi was established in 1991 to protect Mountain gorillas (Gorilla beringei beringei) (photograph 1) and other natural resources. It is within one of the poorest and most densely populated regions of Africa where rural communities depend on natural resources for their livelihood (Plumptre et al. 2004). When local access to Bwindi forest was prohibited under national park status, violent conflict between local communities and park staff arose and, in response, ICD was implemented as a mechanism for conflict reduction and community participation in park management (Blomley et al. 2010; Baker et al. 2011). The first initiatives were collaborative management agreements with specialist resource users from local communities for the collection of minor forest products within harvest zones inside the national park. A series of ICD initiatives followed that included revenue sharing of income from gorilla tourism, crop-raiding mitigation, agricultural development and alternative livelihoods programmes. With this variety of initiatives and success in conflict resolution, ICD at Bwindi evolved to adopt the aim of achieving national park conservation through poverty alleviation. Now, with over 20 years of ICD interventions at Bwindi and a national policy framework that links biodiversity conservation with poverty alleviation, the learning from both Uganda and Bwindi is a valuable resource for the conservation community.

Photograph 1. Mountain gorillas (Gorilla beringei beringei) (credit Julia Baker).

Complimentary Contributor Copy

54

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Chapter Overview In this chapter we seek to review the Ugandan context of conservation-poverty linkages and evaluate ICD at Bwindi as a tool for achieving conservation through poverty alleviation. By identifying lessons learnt from Uganda and Bwindi, our aim is to improve the policy and practice of linking national park conservation with poverty alleviation, particularly to overcome challenges inherent in the ICD approach of reaching the poor and marginalized. Our starting point, in section 3, is to examine historical trends in natural resource management from pre-colonial to the post-independence period when national parks were first established in Uganda. We also assess the legacy of past management regimes on current issues faced by national park managers. In section 4, we analyse the Ugandan policy framework of national park conservation and poverty alleviation and, from this framework, identify the outcomes that national park managers must achieve. In section 5 we introduce our study site of Bwindi Impenetrable National Park, present a conceptual framework of the ICD approach at Bwindi and illustrate the framework by describing the Multiple Use Programme, which has been heralded a success in conflict resolution through collaborative management agreements with local resource user groups. In section 6 we present our first study: a retrospective analysis of interactions between local people and law enforcement rangers as indicators of conflict and local support for conservation at Bwindi. Here we explore drivers of conflict and the factors that engendered local support for the national park during a five-year period (1996-2000) after ICD interventions were first implemented. In section 7 we present our second study: perceptions of local communities regarding governance issues of projects implemented by a major ICD practitioner at Bwindi. Finally, in section 8, we review the lessons learnt from Uganda and Bwindi in the context of forthcoming change to Bwindi’s ICD. We then draw conclusions on the design and implementation of ICD to link national park conservation with poverty alleviation. To achieve conservation goals, reducing unauthorized resource use is often a target of ICD. In this context we define unauthorized resource use as any form of resource harvesting that is not in line with laws or management regulations or conducted without a legal permit. We also define unauthorized resource use not in terms of criminality but as an indicator. Firstly of the different needs and uses of a national park by people: for forest access, to use natural resources and to meet cultural and traditional needs. Secondly of the governance challenges and limitations of national park management to balance people's uses and needs with biodiversity conservation aims.

3. HISTORICAL TRENDS IN NATURAL RESOURCE MANAGEMENT IN UGANDA Pre-Colonial Period In Africa before colonialism, land was generally managed communally. Over time, African societies promulgated rules and regulations on use of natural resources. The rules were precise and codified, although not written down, but were incorporated into the culture (Ochieng Odhiambo 2006; DeGeorges and Reilly 2009).

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

55

Forests In Uganda, communities were socially organized in various ways from kingdoms to chiefdoms and clan systems. The most significant kingdoms were those of Bunyoro, Buganda, Toro, Bunyoro and Busoga. Kingdoms owned the forests, which was managed as either as a communally owned or an open access resource. The former tended to be those that were adjacent to settlements from which people obtained wood and non-wood products for domestic use. Some forests were classified as sacred, in which case an individual or clan was assigned the responsibility of regulating use of its resources with sanctions were imposed for misuse that were learnt by society through stories and folklore. For example, one of the beliefs by Ugandan societies was that, if one went into the forest without reporting the purpose of the visit to the spiritual leader or clan head, that individual would not be able to find their way out and back to the village. Penalties were levied on individuals who broke rules on forest resource use and these ranged from simple ones such as an order to return the removed product to the forest, to severe penalties including ostracism from society (Ggombya-Ssembajjwe, 1995; Turyahabew and Banana 2008; DeGeorges and Reilly 2009).

Wildlife Regarding wildlife, most wildlife resources could be used freely although free access did not imply irresponsible use. Similarly to forests, norms and practices had to be observed and these were passed from generation to generation through strict instruction of the young by the old using stories, taboos, riddles, slogans, tales, proverbs, sayings and song (Osei-Amakye, 1993). Evidence shows that wildlife populations were not as high on community land as in the current protected areas, but were held in check by humans through hunting (DeGeorges and Reilly, 2009). However culturally, many animals, reptiles, birds and fish were venerated and any violation of this taboo attracted supernatural sanctions (Oke, 2007). Among the Baganda for example, folksongs described that the killing of a skink was punishable by not going to heaven. The Baganda also had a tradition of totems, many of which were plants or animals and members of a totem were culturally obliged to protect and defend that species (Nuwagaba and Kiwere in press). There were also community norms for activities such as fishing. Restrictions on equipment such as fishing nets arose from peer-group pressure, social custom and tradition, which amounted to binding law against certain fishing activities (Richardson, 1993).

Summary Therefore before colonialism in Uganda, while natural resources were accessible, there were regulations on use of resources with penalties for misuse. Conservation in the form of protected areas existed through hunting grounds and cultural and spiritual areas that were protected through a set of practices unique to each community. Community leaders were the guardians of natural resources and responsible for regulating access, enforcing rules and adjudicating conflicts associated with access and use.

Complimentary Contributor Copy

56

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Colonial Period The first Europeans to visit Uganda were the British explorers John H. Speke and James Grant, in 1862, during their search for the source of the Nile. As part of the scramble for Africa by European nations, Uganda was declared a British Protectorate in 1894. One of the first acts of the colonial administration was to control natural resource use. The colonial authorities recognized and adopted the systems of land ownership and resource management under the kingdoms, although imposed new regimes to govern natural resources. These included the designation of large areas as crown property where local people were forbidden to enter and collect resources. At this time, large areas of Uganda’s forests were cleared for plantations of crops such as coffee, tea and sugar cane. However, the colonial administration soon realized the need to manage forests, particularly for timber production, and began formal management of Uganda’s forest estate by establishing the Forestry Department in 1917, which was re-named as the Forest Department ten years later (Turyahabwe and Banana, 2008).

Forests The Department was established to direct timber production and manage crown forests, which it did so by entering into a series of agreements on use of forests with each kingdom. The result however included prohibitions on local access to crown reserve forests, which was exacerbated with the first formal forestry policy in Uganda, which was enacted in 1929, led to establishment of more forest reserves. This policy did not address the needs of local people whose livelihoods depended on forests and, in some areas, villagers were displaced from their traditional land without negotiation or compensation. Local access to the forests was allowed only under prescribed conditions although, as these favored the more affluent members of society, communities were gradually alienated from the forests (Turyahabwe and Banana 2008). Conflict between local communities and the colonial authorities arose, which the colonial government responded to by creating local forest reserves that were directly managed by local administrations in an attempt to decentralize forest management. This was the genesis of the two-tier system where larger forests were gazetted as Central Forest Reserves primarily for commercial timber production under control of the government, and smaller forests were gazetted as Local Forest Reserves to cater for local needs. This system was directed by Nicholson (1929) who recognized the dependence of rural communities on forest resources and recommended that supplies of fuel wood, poles and sawn timber be guaranteed by encouraging farmers to grow trees and establish small plantations.

Wildlife Management of wildlife in Uganda during colonial times originated from the culling of elephants. Following Pitman’s (1942) observation on the need to protect local communities from the huge herds of elephant roaming the land, the Elephant Control Department was established to control crop depredation by elephants. Elephant culls were undertaken

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

57

throughout the country with meat from the kills given to local communities and ivory traded to generate revenue. It became the Game Department in 1925 and employed villagers as ‘vermin guards’ to protect the crops of rural communities neighbouring game reserves from elephants and other wildlife species. As observed by Temple-Perkins (1955), by the 1950s “an African guard of the Game Department” was stationed in the vicinity of each community area. This vermin control was a key duty of the department, particularly given the recognition that their efforts to control crop damage by wild animals improved relations with local communities. However, local demand for assistance with controlling crop raiding increased and conflict between local communities and the department arose when local demand was not met. Nonetheless, rural farmers continued to receive assistance with vermin control until the 1980s, when the department’s activities were restricted by the civil war (Uganda Game Department Archives, 1923-1994). In addition to problem animal control, duties of the Game Department included game preservation and reserve management. The Game Ordinance of 1926 mandated the Game Department to regulate hunting and all other forms of wildlife utilization, where trophy hunting by foreigners was permitted although hunting by local communities was restricted. The first wildlife protected areas in the form of Game Reserves and Controlled Hunting Areas were established, and the Ordinance resulted in a strengthening of the game laws and an increase in penalties for illegal activities. Major developments in Uganda’s conservation policies occurred during the 1950s. Under the National Parks Ordinance of 1952, Uganda National Parks (UNP) was established as the government organisation responsible for the management and protection of national parks. Two game reserves were upgraded to national park status: Lake George Game Reserve and Lake Edward Game Reserve were gazetted as the single Queen Elizabeth National Park. Murchison Falls National park was also established and both have remained the largest national parks in Uganda. However their establishment involved the eviction of local people from their traditional land and UNP enforced bans on hunting, natural resource collection and grazing by domestic animals by local people, as well as increasing penalties for offenders. This resulted in widespread exclusion of rural communities from their traditional hunting grounds, burial sites and sacred forests (DeGeorges and Reilly, 2009).

Summary Therefore, natural resource management during Uganda’s colonial era was marked by the transformation of community areas into reserves that were governed under formal laws and regulations. There an attempt to decentralize forest management through a two-tier system of government and locally managed forests. However, establishment of national parks was founded on conservation policies set on protectionist objectives that permeated through not only the laws that were enacted, but also the character of the institutions that were created. This fundamentally changed the conservation landscape redefining the relationship between communities and protected areas as well as legal regimes governing this relationship.

Complimentary Contributor Copy

58

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Independence While the attainment of independence in 1962 was a fundamental milestone in the governance of Uganda, there was no major shift in conservation policy at this time. The colonial legal instruments including conservation laws were codified and published into the laws of Uganda in 1964. Consequently, the philosophy of protectionism as the primary purpose of conservation was inherited at independence and continued to be the hallmark of conservation policy for many years after independence. All natural resources including crown forests became the property of the new independent government of Uganda. However, since the decentralized form of governance in the form of kingdoms and local governments had been maintained, both central and local governments tried to strengthen forest management structures to maximize their benefits from the sector. Whereas it had been agreed during the independence negotiations that all crown forests be returned to local governments, this was on condition that central government was satisfied that local governments had sufficient resources to undertake effective management. By the mid-1960s, no crown forests had been handed over to local governments. They had been left to manage local forest reserves. In 1967, the independence constitution was abrogated and kingdoms were abolished. Local Forest Reserves were turned into Central Forest Reserves and the role of local governments in forest management waned. Forest management became a predominantly central government affair. As was noted by Turyahabwe and Banana, (2004) ‘this change in governance meant that the institutional arrangements that had been instituted by the Local Administrators and forest users to limit entry and harvesting levels lost their legal standing. The decisions regarding forest resource use were entrusted to the Forest Department as the sole agency with powers to regulate the harvesting of forest produce in all Government forest reserves and the use of tree products on public and private land. Thereafter, the Local Administrators were no longer allowed to undertake any forestry work, except maintaining a few village forests, which were not affected by the statutory instrument. The entire Forest Department had little or no downward accountability and limited recourse. This created disinterest in forestry from both local administrators and forest users who viewed forestry as a Government property and no need for its protection’. Developments in the conservation map of Uganda did occur in the 1970s: the AswaLolim Game Reserve, which had been gazzetted in 1959, was degazetted in 1972 and the Kilak Controlled Hunting Area was revoked. However, there were no significant changes to the national conservation policy or agenda because of the breakdown in law and order as President Idi Amin’s government progressively became dictatorial and subsequently collapsed in 1979. The fall of the Amin regime was followed by a period of uncertainty leading up to the elections in 1980 when Milton Obote returned as President. The political uncertainty and instability continued throughout the first half of the 1980s, as the Obote Government was undermined by an insurgency by rebels of the National Resistance Army (NRA). The NRA eventually captured power in 1986 beginning a new phase of Uganda’s conservation discourse. However the political instability post independence constrained both conservation policy-making and practice in the country. In particular, the protectionist approach to conservation that restricted access to natural resources generated widespread public animosity against protected areas. This problem was aggravated by underfunding of conservation activities that resulted in a near collapse of mandated public institutions. A

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

59

combination of all these factors undermined the ecological and legal integrity of protected areas, as major threats such as encroachment, illegal hunting and overharvesting continued unabated.

4. THE LEGACY OF NATURAL RESOURCE MANAGEMENT The historical context of natural resource management can help to understand the attitudinal and behavioral response of local communities towards conservation interventions. In Uganda three factors appear important: resource management systems of local communities, the assistance with controlling crop raiding that authorities gave to villagers and the perception by local communities that national parks are government property.

Resource Management Systems of Local Communities The need for social assessments to inform national park management has long been recognized. However, the more recent acceptance by the conservation community that biodiversity loss and poverty must be addressed as interlocking challenges (Adams et al. 2004) gave additional emphasis to the importance of understanding the socio-economic context of rural communities to develop conservation-poverty alleviation strategies. There is now a growing body of literature on harvesting and use of natural resources by local communities and links between resource use and livelihood security, notably for bushmeat and minor forest products. Community rules, norms and beliefs of resource use is less well studied yet essential to develop conservation interventions that involve resource use or aim to change resource use behaviours. For example, bushmeat hunting in Tanzania is driven by a variety of reasons that include livelihood needs and values but also individual perceptions of what is beneficial, as well as community norms on appropriate and legitimate use of wildlife resources. Therefore, the assumption that poor people hunt for food or income can mask the complexity of this traditional activity and the actors involved, which limits the ability of practitioners seeking to reduce hunting through livelihood improvement initiatives (Bitanyi et al. 2012). A reason why ICD fail is that complex social structure of resource harvesting activities is underestimated or not understood (Bitanyi et al. 2012). Here, by describing resource management systems of Ugandan communities before the colonial period, we provide an insight into the regulations and customs of resource use that included taboos and penalties for misuse. Our intentions are two-fold. Firstly to emphasize the importance of understanding the local socio-economic context of resource use. Secondly to provide a foundation for conservation practitioners to develop a greater understanding of present-day local community norms and individual beliefs on natural resource use, particularly how individuals perceive links between natural resources and their livelihood needs, in order to design and implement conservation-poverty alleviation strategies.

Complimentary Contributor Copy

60

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Assistance with Controlling Crop Raiding and Government Property Box 2. A history of Mountain gorilla - human conflict at Bwindi Impenetrable Forest The first evidence of Mountain gorilla crop raiding around Bwindi is a letter, written during the 1930s, by a prospector working in the Impenetrable Forest to the Chief Game Warden. The prospector described his encounters with gorillas and made the following observation: “the gorillas sometimes raid nearby shambas, but I have never heard of them attacking the natives, and the natives leave them alone except to chase them away from their property” (Uganda Game Department Archives, 1923-1994:1933). Further evidence comes from a report by a game warden of his visit to Bwindi, in 1933, which was then the newly established reserve of Kayonsa. The warden described crop raiding by gorillas and noted that gorillas favoured abandoned cultivated patches: “the Kayonsa gorilla, apparently, is not guilty of frequent shamba-raiding, at least so the natives reassure me. It is true that the gorillas often feed in the vicinity of crops but the attraction is usually the occurrence of various nourishing weeds of exceptional growth which are found on the abandoned cultivated patches” (Uganda Game Department Archives, 1923-1994:1933). The warden also noted conflict issues arising from the presence of gorillas on community land: “the local natives, who can blame them, very naturally object to the proximity of these fearsome beasts, and usually try and drive them away. I am reliably informed that the gorillas are most contemptuous of their efforts, the females and young having been sent off to safety, males only move when it suits them to do so” (Uganda Game Department Archives, 1923-1994:1933). The warden described his trip to see a gorilla group near the forest boundary and the conflict that followed: “when I had seen my fill and was about to retrace my steps, I found at least fifty unauthorised spearmen hanging in the rear, hoping for the opportunity of attacking the gorillas. In fact, I was warned that if I did not personally see this crowd out of the locality, the moment my back was turned they intended going in to spear the male before he could get away from the tree, after which the slaughter of the other four would have been simple. The presence of a European and a misunderstanding would have been their excuse. It shows how easily an unfortunate episode may develop, vide a recent incident in the Belgian Congo, unless all participants in gorilla investigations are absolutely under control” (Uganda Game Department Archives, 1923-1994:1933). The warden also described complaints about gorillas that he received from miners working in Bwindi “prospecting on a systematic scale has taken place in the extreme southerly portion of the forest, but when I was in that neighbourhood at the beginning of November, there were frequent complaints from isolated pairs of natives digging pits, that gorillas were too close to be pleasant” (Uganda Game Department Archives, 1923-1994:1933). Communities around Bwindi did receive assistance with controlling crop raiding from the authorities. One vermin guard was stationed at Bwindi when the forest was under joint management of the Game and Forest Departments (Butynski, 1984) and staff from both departments regularly assisted farmers by scare-shooting when wild animals foraged within agricultural land. Game guards, in particular, would respond when large animals, such as elephants, entered community land. The guards would also kill smaller animals that frequently raided crop and livestock including baboons and bushpigs (Namara, 2000). Vermin control remained a duty of law enforcement rangers after Bwindi was designated a national park. Farmers would request assistance when rangers passed their fields while patrolling the national park boundary or would travel to the ranger outpost to request assistance. Rangers would employ scare shooting for elephants and monkeys, and help farmers to chase gorillas and duikers into the forest by shouting and beating drums. However problem animal control was a secondary duty for rangers after law enforcement and gradually phased out when ICD initiatives for problem animal control, which included Mountain gorillas, were introduced (Baker 2004).

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

61

In Uganda crop raiding by wild animals has been documented since the early 1900s. Historical records include crop raiding by Mountain gorillas around Bwindi Impenetrable Forest during the 1930s (Box 2). Here we describe how colonial authorities employed local people as vermin guards to control crop-raiding by wild animals, and how central management of reserves led to local populations viewing the reserves as government property. Crop raiding is currently a significant cause of conflict between local communities and national park authorities in Uganda, although the conflict often arises when local communities perceive that the authorities have failed to assist them to control crop raiding (Baker 2004). The legacy of the colonial authorities providing assistance to control crop raiding may have contributed towards this conflict, as local communities still speak of the assistance they used to receive (Baker 2004). Furthermore, the conflict could be compounded by the view of local communities that national parks are government property and wild animals in the parks are the government’s responsibility. Learning from previous management regimes is therefore important for designing interventions to reduce crop raiding. With Uganda, the Game Department documented how assistance with controlling crop raiding improved their relations with local communities, but also documented the rise in conflict when an increase in local demand for assistance was not met. If interventions to reduce crop raiding are to be implemented, managing local expectations and involving communities in the design and implementation of the intervention are therefore important. Furthermore, if local communities believe that a national park is owned by the government and that the government is responsible for its management, ICD practitioners must consider how best to involve communities in managing the park so that they do feel a sense of ownership and, when ownership has been achieved, a voluntary commitment to safeguard the park. One example of an approach to engender a sense of ownership and encourage individuals to adopt the role of ‘national park guardians’ is the Multiple Use Programme at Bwindi Impenetrable National park.

5. UGANDA’S POLICY FRAMEWORK FOR CONSERVATION AND POVERTY ALLEVIATION Contemporary national policy framework for conservation and poverty alleviation in Uganda can be categorised into four phases: the colonial phase with resource management regimes that excluded communities from traditional lands, focused on revenue generating and included problem animal control the post-independence phase up to 1992 where the establishment of national parks further alienated communities for their lands the post-UNCED phase up to 2010 where poverty was identified as a key driver of environmental degradation the National Development Plan (2010) to present day with emphasis on wealth creation as the vehicle for attaining poverty eradication and conservation policy objectives

Complimentary Contributor Copy

62

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Our last section described the policies and natural resource management regimes of the colonial and post-independence phases. In this section we review developments by the global environmental movement to link conservation and poverty alleviation and then analyze Uganda’s response to the movement to link conservation and poverty alleviation in both policy and practice.

Beyond Environmental Protection: The Global Discourse on Conservation and Poverty Alleviation In 1972 the international community adopted the Stockholm Declaration and a plan of action for the environment outlining an ambitious agenda linking conservation and development (See UN GA RES 37/7 (1982)). The term poverty was not reflected in the final documents of the Conference, although the term “development” was used in the Declaration and in the Plan, and also in the World Charter for Nature that was adopted by the United Nations General Assembly in 1982. The outcomes of Stockholm and the World Charter reflected a growing realization that preservationist strategies for natural resource protection were ineffective and that conservation needed to be reconfigured to achieve objectives of sustainable development. There was then a global shift in the management and conservation of protected areas. The lexicon of poverty entered the conservation discourse in the 1980s during the work of the World Commission on Environment and Development (WCED), and was used extensively in its final report entitled “Our Common Future.” In its report, the Bundtland Commission drew a clear nexus between conservation and poverty noting that a world in which poverty is endemic will always be prone to ecological and other catastrophes. By the United Nations Conference on Environment and Development (UNCED), which was convened in Rio de Janerio, Brazil, in 1992 to respond to the Brundtland Commission, there was already growing consensus that conservation policy should address issues of poverty. All the three key outcomes of the UNCED emphasized the centrality of poverty eradication as an overriding goal in the design of conservation policies at all levels. These outcomes were a political statement commonly referred to as the “Rio Declaration”, an action plan termed “Agenda 21” and a legally binding agreement on the conservation of biological diversity. Parties to the Convention acknowledged that the extent to which developing countries will implement their commitments to conserve biological diversity ‘will depend on the effective implementation by developed countries of their commitments under this Convention related to financial resources and transfer of technology and will take fully into account the fact that economic and social development and eradication of poverty are the first and overriding priorities of the developing countries’. In effect, the Rio Declaration and subsequent global processes placed poverty firmly on the agenda of conservation and emphasised the role of protected areas governance to achieve this agenda. New strategies such as “collaborative forest management”, “sustainable use of forests” and “equitable sharing of resources” emerged (CBD, 1993). These strategies were later emphasized and strengthened by the 2003 fifth world park’s congress held in Durban. Uganda became a signatory to the CBD on 12th June 1992 and, being part of the 2003 World Park’s Congress, was obliged to involve local people in protected area management and adopt the mandate of conservation through poverty alleviation.

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

63

The Post-UNCED Conservation Policy and Poverty Eradication Discourse in Uganda Deliberate efforts to link conservation, protected areas and poverty alleviation in Uganda are a recent phenomenon. These efforts began in 1986 when the National Resistance Movement (NRM) and its military outfit, the National Resistance Army (NRA), took power and formed a new government. The conservation consciousness of the NRM leadership was evidence in its blue print document – the Ten Point Programme – that made reference to ‘the continued destruction of the environment’ and ’the profitless extraction and export of exhaustible natural resources’. The ministry of tourism and wildlife was among the first ministries to be created when the new government was formed and, in 1987, a specific ministry responsible for environmental protection was created. In 1989, forestry was transferred from the ministry of agriculture and combined with the ministry for environment protection. Then in 1992, a realigned ministry dealing with energy, minerals and environmental protection was established. During this period until the promulgation of a new constitution in 1995, there were significant policy developments in terms of protected areas governance. The most significant of these developments were: (i) the adoption of the National Environment Action for Uganda, 1994 (ii) the promulgation of the National Environment Management Policy, 1994 (iii) the enactment of the National Environment Statute (iv) the promulgation of a new constitution, 1996 (v) the enactment of the Uganda Wildlife Statute, 1996 Since these instruments, the commitment to simultaneously achieve the objectives of conservation by securing the ecological and legal integrity of protected areas, and of poverty eradication has remained the hallmark of Uganda’s national policy framework. The dominant policy narrative on the environment and poverty nexus was defined in the National Environment Action Plan 1994 in the following terms: The key link between poverty and environment is that poverty affects people’s ability to manage their environment sustainably. As they lack resources and appropriate technologies, many farmers must resort to cultivating steep slopes, erosion-prone hill sides, semi-arid lands or encroach on the protected areas in order to meet their various demands. In short, poverty compels them to destroy those very resources that are necessary to relieve them of hunger, disease, and further poverty.

This narrative makes clear that poverty is a driver of environmental degradation and a threat to protected areas. This theme continued through the various policy instruments including the National Environment Management Policy. This policy recognized, inter alia, that poverty and the degradation of natural resources and the environmental were so intertwined that they required an integrated approach to address them. It observed that the need to reorient national and local efforts to address environmental problems in a more comprehensive and integrated manner ’will constitute the fundamental basis for achieving overall policy goal of sustainable socio-economic development which maintains and

Complimentary Contributor Copy

64

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

enhances environmental quality and resource productivity to meet the needs of present and future generations’. Although the term ’poverty’ is used in the policy only three times, it is clear from its principles and strategies that poverty eradication is both the means and the ultimate objective of natural resource management. For example, the policy sets out the objective of population, health and human settlement as ‘to manage population growth, settlements, distribution and health in such a way as to match people and resources in an economically, socially acceptable and environmentally sound manner’. Strategies to achieve this objective include actions to ‘promote income generation programs which aim at the alleviation of poverty especially among women and lower income groups’ and ‘facilitate women’s participation in population and environment decision making, resource ownership and management, as well as improve their access to inputs including better access to credit’. The critical relationship between conservation and poverty alleviation is further addressed under sections of the policy that detail the two key management regimes of protected areas: forestry and wildlife. Under forest conservation and management, the policy provides that local community involvement in the planning and management of protected areas and in the sharing of benefits derived from protected areas is crucial for conserving forest resources. Three of the strategies outlined to achieve these objectives of sustainable forestry management emphasize that community participation and revenue sharing - and the governance required to achieve both - are central tenets of conservation policy and ethic (Box 3). Box 3. Strategies to achieve sustainable forestry management in the National Environment Management Policy of Uganda 1994 (ii) Revise and strengthen the Forest Act with particular regard to gazetting and degazetting, collective responsibility in management, revenue sharing and local community participation in PA management (iii) Improve local capacity to manage protected and gazetted forest reserves by encouraging people's participation in forest planning and management (xvii) Enhance local community participation in the management of protected areas, where feasible, through the development of Forest Management Advisory Committees, cooperative co-management agreements, and parish and sub-county workshops, and provide more direct benefits to local communities from protected area activities including the return of a percentage of revenue to them.

Similar principles and strategies are captured in the policy section on wildlife conservation and management (Box 4). Here the policy provides that ’the involvement of local communities in the planning and management of protected areas and in the sharing of benefits derived from these areas is crucial for the conservation of wildlife resources’. This illustrates that good governance is the foundation for efforts to link protected area conservation with poverty alleviation and that the output of ‘good governance’ is important to measure when evaluating success of these efforts. This policy therefore clearly establishes practical strategies for linking protected area conservation with poverty alleviation. However, two of the key policy instruments adopted after UNCED (the 1995 Constitution and the Uganda Wildlife Statute) did not include

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

65

poverty alleviation as a central tenet of protected area governance. Nevertheless, the 1995 Constitution defines a set of guiding principles whereby management of Uganda’s protected areas is to benefit local people. In addition, the Uganda Wildlife Statute established a system of ‘wildlife use rights’ that could be a basis for designing protected area programmes to ensure mutuality between protected area management and poverty alleviation. Box 4. Strategies to achieve wildlife conservation in the National Environment Management Policy of Uganda 1994 (iii) Develop a policy framework and guidelines for the identification and management of buffer zones and buffer areas in and around protected areas to help reduce conflicts between multiple uses and users (e.g., livestock and wildlife) (iv) Establish a mechanism for collaboration between protected area management and the neighbouring communities in order to resolve potential conflicts through the involvement of local people in the planning, management and decision making process, and ensure that a portion of benefits from the protected area system is offered to the local communities (v) Enhance local community participation in the management of protected areas through the development of Parks Management Advisory Committees, parish and sub-county workshops, etc., and provide more direct benefits to local communities from protected areas activities including the return of a percentage of revenue to them (vi) Establish a mechanism for collaboration between protected area management and the neighbouring communities in order to resolve potential conflicts through the involvement of local people in the planning, management and decision making process, and ensure that a portion of benefits from the protected area system is offered to the local communities

A number of conclusions can be made from the immediate post-UNCED national policy instruments for governing protected areas. Firstly, by the mid-1990s, the relationship between protected area management and poverty alleviation was widely accepted in policy-making circles in Uganda. Secondly, the dominant policy narrative in these instruments, as well as the practice of the mandated institutions, identified poverty as a major threat to protected area conservation. Consequently, many conservation interventions during this period relied on law enforcement and legal sanctions to address problems confronting protected areas. Thirdly, the mandated agencies were configured explicitly as protection or conservation agencies and were not properly retooled to serve the goal of poverty alleviation. Finally, poverty alleviation was then seen not as a goal but a vehicle to achieve objectives of conservation. This policy discourse on protected area management and poverty alleviation remained on parallel tracks until the late 1990s when the Poverty Eradication Action Plan process (PEAP) commenced.

From The Poverty Eradication Action Plan to the National Development Plan There was widespread perception that the promulgation of written policies and enactment of legislation providing incentives for public participation and benefit sharing would fundamentally change the conservation landscape in Uganda. However, the reforms of the

Complimentary Contributor Copy

66

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

early 1990s did not result into substantive programmes until 1997 when the Poverty Eradication Action Plan (PEAP) was adopted. For over a decade (1997-2010), PEAP acted as Uganda’s national development policy framework and medium term planning tool. PEAP put the poverty alleviation agenda at the forefront of development planning and budget policy in Uganda. By bringing together the conservation community, planning and budget policy makers, political leaders, the civil society and a broad range of stakeholders, the PEAP process itself enhanced the policy discourse on the environment, protected areas and poverty nexus. Several key policies were promulgated during this period. In particular, the Plan for Modernization of Agriculture (PMA) sought to channel appropriate investments in agriculture and natural resource management as a strategy to eradicate rural poverty, and the National Forestry Policy outlined a comprehensive national agenda with emphasis on harnessing forestry resources to contribute to poverty eradication.

From Policy to Practice: The ICD Approach There is compelling evidence that, over the last two decades, a comprehensive policy regime that provides the framework for achieving convergence between conservation and poverty alleviation objectives has been established in Uganda. These polices have been complemented by laws and institutions with wide ranging mandates to take actions to achieve the convergence. No matter the shortcomings, the policy framework that evolved over the years provides a framework for targeted interventions to simultaneously pursue the objectives of protected area conservation and poverty alleviation. This includes the aim to redress the inequitable distribution of the costs and benefits of conservation. The ICD approach has been adopted at many protected areas in Uganda. It was based on the premise that communities living around protected areas incur substantial costs on account of restricted access to natural resources and alienation of traditional lands. Our next section reviews the ICD approach at Bwindi Impenetrable National park in southwest Uganda.

6. BWINDI IMPENETRABLE NATIONAL PARK Located in south-west Uganda (Figure 1), Bwindi covers 330.8 km2 of dense forest with a rugged topography of narrow valleys and steep hills and elevations ranging from 1200 m to 2600 m (Plumptre et al. 2004) (Photographs 2 and 3). A small section of the western park boundary borders the Democratic Republic of Congo and the remaining boundary is bordered by 21 densely populated parishes. At gazettement, average population densities were 125 people/km2 in central and northern areas, 256 people/km2 in eastern areas, and 275 people/km2 in southern and western areas (UBOS 1991). This is one of the poorest and most densely populated regions of Africa, where rural communities depend on natural resources for their livelihood (Plumptre et al. 2004).

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

Photograph 2. Forest-community boundary of Bwindi Impenetrable National Park (photo credit Julia Baker).

Photograph 3. Bwindi Impenetrable National Park (photo credit Julia Baker).

Complimentary Contributor Copy

67

68

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Local People and Bwindi Forest Villagers neighbouring Bwindi rely on agriculture and forest products for their livelihoods. Farming is mainly for subsistence but provides an income from cash crops that include tea and from sales of surplus subsistence crops at local markets (Plumptre et al. 2004). Although described as rural subsistence-based communities, the villagers seek multiple sources of income and, before gazettement, Bwindi forest provided most of their non-farming income when traders employed villagers as labourers for pit sawing and mining within Bwindi and for smuggling cattle and other goods across international borders through Bwindi. Trails through Bwindi also provided access to markets for villagers to sell crops, crafts or forest products. The sale of forest products provided both an income for the seller and a source of forest products for villagers, particularly wood for furniture. Forest products were used by crafts people for their trade. For example blacksmiths, a small but important group of specialists for the farming community, used the forest tree species Polyscias fulva to construct bellows for producing farming tools. Community groups also relied on forest resources, such as the stretcher-bearer societies. These societies transport sick (or dead) people within the Rukiga highlands surrounding Bwindi. The societies are well-organised and receive financial payments from villagers on a monthly basis to cover the cost of food for journeys and for buying new stretchers. Stretchers last 2-4 years depending on the materials used and, before gazettement of Bwindi, were woven from leaf stems or plants from the forest. In addition to providing direct and indirect income sources, Bwindi provided subsistence resources for villagers including forest products such as firewood and beanstakes, and resources for specialist forest users including minor (non-timber) forest products of medicinal plants, basketry materials and food. Food included honey, edible plants and bushmeat. Although bushmeat hunting was an important cultural tradition, bushmeat was primarily sought for domestic consumption and provided only a modest income for local hunters (Tukahirwa & Pomeroy 1993; Cunningham 1996).

Conservation Importance Bwindi contains one of the two remaining small populations of the critically endangered mountain gorilla. The forest is the only site in East Africa with a continuous forest cover over an altitudinal range of 1190-2607m and, as a Pleistocene refugium, the highest biodiversity site in East Africa for various species including rare and endemic species (Butynski 1984; Kingdom, 1990; Howard, 1991; Hamilton et al. 2000; McNeilage et al. 2006).

History of Management Bwindi Impenetrable National park comprises two blocks of forest connected by a small corridor with approximately a 115km long boundary. It was first gazetted as a forest reserve by the colonial government in 1932. In 1961 Bwindi became a game sanctuary under joint management of the Forestry and Game Departments until 1991, when it was gazetted as a national park and became under management of the Ugandan National Parks (later renamed

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

69

as the Uganda Wildlife Authority (UWA)). Mountain gorillas are the flagship species for conservation efforts at Bwindi and, before establishment of the national park, threats to gorillas included habitat loss from timber pit sawing and mineral prospecting, and death or injury from bushmeat snares. Commercial activities of pit sawing and mining had been conducted in Bwindi since the 1930s. At gazettement pit sawing was the most prevalent human activity occurring throughout Bwindi, whereas mining was concentrated in the centre of the forest in or near water. At the time of gazettement, Bwindi was an island surrounded by intensively farmed land that in many places extended to the boundary (Butynski 1984; Howard, 1991; Cunningham, 1996).

Conflict Under national park status, intensive ranger patrols were implemented to enforce the prohibition on local access to Bwindi. Conflict between local communities and conservation authorities then arose and included violent incidents where rangers were attacked and fires within the national park were started deliberately. The conflict arose because of various factors including loss of subsistence resources that local people gained from the forest (Wild & Mutebi 1996; Hamilton et al. 2000; Blomley et al. 2010). However the violent conflict was primarily triggered by rangers arresting pit sawyers or miners. The conflict therefore arose because of the loss of financial benefits to villagers that these trades generated, which included employment for villagers, income for traders and timber and mineral markets for other villagers to sell crops or crafts (Baker et al. 2011).

ICD at Bwindi Forest: A Conceptual Framework In response to the conflict, ICD was adopted as a mechanism for conflict reduction and community participation in park management. The initiatives comprised linking and delinking strategies. Linking strategies aimed to increase local support for conservation by generating benefits from the national park. The first strategies were collaborative management agreements with specialist resource users whereby zones inside Bwindi were established for beekeeping (1991 and 1992) and collection of herbal medicines and basketry materials (1994) (Figure 2). Tourism focused on viewing mountain gorillas began in western areas in 1993 and brought employment and trade opportunities. Tourism revenue sharing and a trust fund were established in 1994 to support community projects, including the construction of schools and health clinics. By contrast, delinking strategies aimed to decrease pressure on forest resources by providing alternative incomes and resources. These strategies included agricultural extension programs to reduce demands for forest land by increasing productivity on existing farmland, and on-farm substitution of forest products, including firewood (Blomley et al. 2010). Many of Bwindi’s ICD initiatives were designed to reduce conflict and have been designed and implemented with the participation of local communities (McNeilage & Robbins 2006). The ICD has since developed to seek to achieve both conservation and poverty alleviation goals.

Complimentary Contributor Copy

70

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Figure 2. Harvest zones of Bwindi Impenetrable National Park.

Multiple Use Programme: Harvesting Resources in Bwindi Impenetrable National Park Forest resource use by local people around Bwindi is as old as mankind that has lived there. For centuries Bwindi forest has been a source of livelihood for the local people. The forest was a source of protein from bush meat and fish for local people. It was also used for extraction of plant resources for food, basket weaving, medicinal purposes and house construction.

Multiple Use The Multiple Use Programme (MUP) was established at Bwindi on the premise that giving local people access to the national park to harvest minor forest products for livelihood subsistence would improve park-community relations and would stop or reduce unauthorised resource use. Through collaborative management agreements with specialist resource users, MUP was also a strategy to involve local communities in the management and protection of the national park and re-establish a sense of forest ownership. The term ‘multiple use’ initially referred to multiple land-uses of Bwindi, i.e. biodiversity conservation, tourism,

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

71

research and low impact forest resource use. However as MUP developed, the term came to mean low impact forest resource use only (Cunningham, 1996; Wild, 2001)

A Pilot Scheme Following an extensive review of how local people neighbouring Bwindi harvested and utilised forest resources, three resources were selected by conservation authorities on the basis that the harvesting activities were of low impact on the national park. The resources were plants for basket weaving and medicinal use and beekeeping for honey collection. Periphery areas of the national park for the resource harvesting (called harvest zones) were established in collaboration with local communities. The zones extend a maximum distance of 2km from the national park boundary into the forest interior and 20% of the total area of the national park was designated as harvest zones (Cunningham, 1996; Wild & Mutebi, 1996; Wild 2001; Bitariho et al., 2006). Harvest zones started as a pilot scheme shortly after gazettement of Bwindi in 1991 for beekeepers in eastern areas. Following success of the pilot, in 1994 MUP was extended to parishes of Mpungu, Rutungunda and Nteko for plant collection and Kitojo, Nyamabare, Kashasha, Nshanjare and Byamihanda parishes for beekeeping. The programme was implemented by UWA with support from ICD organisations. Authorized resource users with given identity cards and Memorandums of Understanding (MoU) were established between UWA and each MUP parish. The MoU detailed the harvesting activities and quotas and defined the role of authorized resource users as forest guardians that included assisting rangers with law enforcement activities. In 1994 there were 187 authorized plant harvesters and 378 beekeepers. In 1999 MUP expanded to include 144 authorized plant harvesters in three additional zones for plant harvesting (Karangara, Masya-Kifunjo (now called southernward) and Remera). This resulted in a total of 709 authorized resource users under MUP at Bwindi (Wild & Mutebi, 1996; ITFC, 1999; Wild 2001; Bitariho et al., 2006 and 2006b).

Success? Several reviews of the MUP have been carried out with an overall aim of involving more local people in park management e.g. Bensted-Smith et al. (1995); Davey et al. (2001) and Bitariho et al. (2004). All the reviews recommended the expansion of the MUP to include other areas not benefitting from the programme. However, the expansion of the programme to other parishes has been limited by the 20% quota allocated for multiple use zones. The 13 multiple use zones (MUZs) at the BINP park periphery had already covered the 20% quota allocated for the MUP. In 2001, a new management plan for Bwindi recommended the expansion of gorilla tracking to other new forest areas (apart from Buhoma) and this included Nteko multiple use zone (UWA, 2001). Since it was deemed that both tourism and multiple use zones could not exist together, a recommendation was made in 2002 where Nteko MUZs was replaced with a tourism zone. This therefore reduced the original MUZs from 13 to 12 zones, see figure 2(UWA, 2001). The total number of registered authorized resource users in Bwindi also reduced to 667 people.

Complimentary Contributor Copy

72

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Implications of Changes Presently, a new management plan for Bwindi is being formulated (2012 to 2022) in which the MUZs will be reduced to a further 10 zones when the Kitojo and Mpungu MUZs are replaced with gorilla tourism zones as recommended by the draft plan (Figure 3). This would further reduce the number of authorized resource users around Bwindi to a further 478 people contributing to a percentage reduction of resource users in BINP to about 33%. Furthermore, the new management plan is proposing to reduce the maximum distance of the MUZs from the forest edge into the interior to 1 km from the original 2km see figure 3 (UWA, 2012). All these events in the Bwindi’s MUP are recipes for future conflicts between the local people and park managers. It can be argued that tourism programmes in Bwindi provide income to the local people than the MUP. However evidence around BINP and elsewhere suggests that tourism activities tend to benefit only the elite and outsiders than the local resident poor people who lack the skills, knowledge and resources to tap from the tourism activities (Arnold & Perez, 2001; Newton, 2007; Sandbrok, 2009; Blomley,et al., 2010). In order for Bwindi’s park management programs to benefit the poorest people like the elite, the multiple use zones should be allowed to exist alongside the tourism zones. Indeed this was the case in Kitojo a new tourism zone before park managers suggested otherwise without any scientific data to back such action.

N

W

E

BINP Management Zones

S

KEY

Roads 0 All We athe r, Loose Surfac e Dry W ea the r, Loose Surfac e Motorable Tr ack Footpath Patrol / Res earc h Tr ail Public Footpath in Pa rk Touri st Tr ail Defunc t Tr ail Boundar y Sm all rive rs Sea sonal ri vers Big riv ers Wil de rness zone Resourc e use zone Touri sm zone

5

0

5

10 Kilometers

Figure 3. Proposed Bwindi park management zones (including new multiple use zones also known as resource use zones) (source UWA, 2012) .

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

73

Like most collaborative forest management programmes elsewhere, Bwindi’s multiple use programme is a result of conflicts between natural resource managers and adjacent local communities. Local people depend on Bwindi forest for their livelihoods and see the forest as a source of insurance against environmental catastrophes such as droughts and floods. When local people’s livelihoods are threatened and limited by park managers, there is bound to be conflicts. Although the multiple use programme was initiated to help mitigate such conflicts, the recent changes in the MUP might initiate new conflicts between park managers and the local people. There is need to find a balance between natural resources conservation and the need for local livelihoods in Bwindi forest. Although some elite local people are bound to benefit from other park programmes such as tourism, the poorest local people such as the Batwa do not have resources and skills to tap from the tourism potentials provided by replacing the multiple use zones with tourism zones. Therefore, Bwindi’s newly proposed management plan should not exacerbate the conditions of the poorest local people by further replacing the multiple use zones with tourism zones. The two zones could exist together as a compromise since there is no scientific data to the contrary.

Summary: ICD at Bwindi

Photograph 4. Homestead neighbouring Bwindi Impenetrable National Park (photo credit Julia Baker).

Bwindi’s mountain gorilla population occurs within one of the poorest and most densely populated regions of Africa. This creates major challenges for Uganda to conserve gorillas and ensure that conservation contributes towards local livelihood improvements (photograph 4). After ICD was adopted at Bwindi and MUP was implemented, conflict declined and local attitudes toward the park became more positive (Wild & Mutebi 1996; Hamilton et al. 2000;

Complimentary Contributor Copy

74

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

McNeilage & Robbins 2006). The MUP was heralded an example of successful collaborative management between conservation authorities and local communities that involved local people in national park management and re-established their sense of forest ownership. A recent review of ICD at Bwindi found that ICD was important for improving parkcommunity relations but had several flaws: it tended to benefit wealthier community members rather than the poorer households assumed to be undertaking illegal activities and had little impact on reducing threats posed by illegal activities. While law enforcement has reduced illegal activities substantially since gazettement (Blomley et al. 2010), threats to gorillas remain. For example, gorillas were killed by local people illegally hunting for bushmeat inside the park in 1994 and in 2011 (Amooti, 1995; IGCP 2011). Therefore, ICD at Bwindi has improved park-community relations although has not been effective in linking conservation and poverty alleviation.

7. INTERACTIONS BETWEEN LOCAL COMMUNITIES AND LAW ENFORCEMENT RANGERS AS INDICATORS OF CONFLICT AND LOCAL SUPPORT FOR CONSERVATION AT BWINDI IMPENETRABLE NATIONAL PARK, UGANDA We undertook a historical analysis to examine responses that local community members made to law enforcement rangers at Bwindi Impenetrable National park from 1996 to 2000, a five-year period after MUP had been implemented. Our aims were to assess causes of conflict between local communities and park staff and identify factors that engendered local communities to support the national park. Our objectives were to determine the types of positive and types of negative responses by communities to rangers, and the factors that best explained whether communities responded positively or negatively towards rangers. The analysis was based on the hypothesis that individuals directly benefitting from MUP responded more positively to rangers than individuals not directly benefiting. Individuals directly benefitting from MUP were defined as authorised resource users of the MUP.

Methods Retrieval and Verification of Law Enforcement Reports We retrieved law enforcement patrol reports for Bwindi from 1986 to 2000 from the national park headquarters and ranger outposts around the park. The reports were handwritten accounts by rangers of their encounters with unauthorised resource use and wildlife while on patrol. We verified recordings by rangers to validate data within the reports (Baker 2004; Baker et al. 2011). Community Response Data From 1996, rangers recorded their interactions with, and observations of, members of local communities in their patrol reports. These recordings came under the heading of ‘community response’ and consisted of descriptive notes detailing conversations with community members and general observations made by rangers on the attitude of local

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

75

communities towards the national park. All ranger-community interactions were made outside the national park when rangers patrolled the national park boundary or when rangers returned to their outpost after a patrol through community land.

Validation of Community Response Data From 1996 to 2000, rangers recorded 445 responses by local communities from a total of 1288 patrol days. We validated rangers’ recording of community responses using the same methods that we used to validate rangers’ recording of law enforcement efforts (Baker 2004; Baker et al. 2011). Our validation showed that most (87%) ranger recordings of their interactions with local communities were assigned the same categories that were assigned from our recordings made while accompanying the patrols. We therefore considered that rangers’ recordings were representative of ranger-community interactions and this validation permits confidence in the accuracy of rangers’ recording of community response. Constructing Typologies of Community Response Data We developed three typologies of the community response data: i. ii. iii.

the type of response (i.e. whether positive or negative) the community member making the response the location of the response

I. Type of Response From descriptions that rangers made in the patrol reports of their conversations and interactions with community members, we developed a five-point Likert scale that ranked from very negative to very positive on types of responses by communities to the rangers. We presented the scale to law enforcement rangers, community conservation rangers, national park wardens and staff of ICD organisations for their verification of the positive and negative responses. We then held focus group discussions with community conservation rangers and local community leaders to further refine positive and negative responses. Obtaining the local community perceptive was also to minimise bias in the data, as descriptions of community responses in the patrol reports were only from the rangers’ perspective and no documented evidence by local communities of the interactions was available for this study. The focus group discussions confirmed the difference evident from the rangers’ descriptions in the level of conflict between complaints about crop raiding and requests for assistance to control crop raiding animals. Villagers complaining to rangers about crop raiding would typically just complain, whereas those asking rangers for assistance would often become aggressive, particularly when their requests were not met. The outcome from the focus group discussions was to categorise complaints about crop raiding as negative but requests for assistance to control crop raiding animals as very negative. Previous evaluations of ICD at Bwindi have used actions by local communities regarding fire in the national park as indicators of conflict and support for conservation. For example, the deliberately started forest fires during gazettement of Bwindi demonstrated resentment of the national park whereas assistance by local communities to help park staff with fire control indicated their willingness to become engaged with protecting the National park (Blomley et al. 2012; Baker et al. 2011). In patrol reports from 1996 to 2000, rangers recorded two

Complimentary Contributor Copy

76

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

incidents when communities reported fire to rangers, both in 1999 by villagers around the centre. Rangers also recorded 23 incidents when they received assistance by local communities with fire control, which were in 1999 and 2000. Reporting fire and assistance with fire control were categorised as very positive responses. However, discussions with park staff and local community leaders revealed that some forest fires were deliberately started by villagers who then assisted rangers with fire control to receive a reward. Therefore, the true nature of responses concerning fire during this period of Bwindi’s history was difficult to determine. The outcome of focus group discussions was to omit fire responses from the statistical analysis although to include the responses as part of the interpretation. The final definitions on types of community responses comprised a five-point Likert scale that ranked from very negative to very positive (Table 1). Table 1. The Likert scale for the type of response by communities to law enforcement rangers in Bwindi from 1996 to 2000 Type of Response Very negative

Negative

Neutral Positive Very positive

Definition Refuse to assist rangers investigating unauthorised resource use Refuse to assist rangers with the trial of arrested offenders Alert offenders inside the national park to rangers on patrol Aggressive request of compensation, vermin guards or land purchase because of crop raiding by wild animals Complain to rangers about crop raiding animals Complain to rangers about living adjacent to the national park Complain to rangers about community projects of ICD Report suspected problems in the forest to rangers, e.g. dead animals from Enquire about national park related issues Positive comments about National park and/or conservation Appreciation for rangers’ assistance with problem animal control Assist rangers investigating unauthorised resource use Report unauthorised resource use to rangers on patrol Report unauthorised resource use to rangers at the outpost

II. Community member From rangers’ notes in the patrol reports, we identified four types of community members who were made responses to rangers: villagers; village and parish councillors (hereafter referred to as local councillors); authorised resource user (defined as individuals directly benefitting from MUP) and the Batwa (hunter-gatherer communities). A concern arose about a separate category for authorised resource user because authorised resource users are members of the local community and therefore also villagers. Although many rangers recorded whether their interaction was with a villager or authorised resource user, some might have recorded 'villagers’ if they did not know whether an individual was an authorised resource user, or knew but just recorded ‘villagers’. Discussions with park staff and local community leaders revealed that rangers knew the authorised resource users (many rangers themselves being local community members) and that these individuals were referred to as ‘authorised resource user’ because of the status that this gave them in their community. This permitted confidence in the distinction that rangers made

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

77

between villagers and authorised resource users, which enabled comparisons of the type of response (whether positive or negative) by individuals directly benefiting from MUP and individuals not directly benefitting MUP. There were two responses by the Batwa to rangers, which were both around the south of Bwindi. A Batwa man reported illegal pit sawing in the national park to rangers on patrol in 1996, which ranked very positive. In contrast, a group of Batwa men working in fields adjacent to the national park boundary alerted offenders inside the forest to an approaching patrol in 1998, which ranked very negative. These responses were noted although excluded from the analysis because of the small number of responses by the Batwa that rangers recorded.

III. Area of the response In some patrol reports, rangers recorded the local name (topoymn) of the specific area of the community response. However there were patrol reports where only topoymns of the whole patrol were recorded with no details on where the community response occurred. For consistency in the data, we assigned the area of community response to one of the five patrol areas of Bwindi: north, centre, south, east and west. These patrol areas had been developed from geo-referencing topoymns (as described in Baker, 2004 and Baker et al. 2011). Most (92%) community responses were recorded in one patrol area only. The remaining community responses comprised patrols where rangers crossed from one patrol area to another and no specific location of the community response was recorded. Here we assigned the community response to the patrol area where rangers spent most patrol time on the national park boundary, as this was the most likely location of the community response.

Analysis We conducted a historical, correlational analysis to examine the community members who responded negatively and who responded positively to rangers, and to examine temporal and geographical patterns of the negative and positive responses. We analysed data by monthly totals and summed the number of community responses per patrol day for each of the five patrol areas of Bwindi per calendar month per year (1996-2000 monthly totals for all areas; n = 141). 1. Was a particular area or community member significantly associated with positive or negative responses to rangers? We first aimed to identify whether a particular area or community member was associated with positive or negative responses to rangers. We used the Chi Square test to examine associations between positive and negative response and area of Bwindi, and between positive and negative response and community member. The low number of responses by authorised resources users did not permit statistical analysis, so we grouped responses by authorised resource users into three categories of negative (negative and very negative), neutral and positive (positive and very positive) in order to conduct the association test.

Complimentary Contributor Copy

78

J. Baker, R. Bitariho, A. Gordon-Maclean et al. 2. Did either location or type of community member significantly influence whether rangers received a positive or negative response?

Our second analysis was to determine the significance of area and community member to the type of response that rangers received. The data comprised the number of community responses per month, which we analysed by log linear analysis under the assumption of a Poisson distribution, using the hierarchical approach and specifying a log link function. The west area and the Batwa were omitted from the log linear analysis because of the small numbers of responses. A three-way (4x3x5) contingency table was constructed with the factors of area (north, centre, south, east), community member (villager, local councillor, authorised resource user) and type of response from very negative to very positive. When the final model was generated, we examined standardised lambda values of the interaction terms to determine patterns in community response that best explained the model. 3. Did the number of positive responses that rangers received significantly differ between locations around Bwindi, year (between 1996 and 2000) or season? Our third analysis was to determine whether the amount of positive responses that rangers received differed between area, year or season. The number of positive responses per month was expressed as a proportion of the total number of community responses. This formed the dependent variable for the analyses, which were undertaken using the nonparametric tests of Kruskal-Wallis, Mann Whitney U and Spearman’s rank correlation. Comparisons were undertaken of the mean proportion of positive community response between the five patrol areas of Bwindi, years from 1996 to 2000, months of the year, months of the rainy and dry seasons and months of the different farming seasons (planting, growing, harvesting) around Bwindi. Months of the rainy and dry seasons and farming seasons were developed based on discussions with community conservation rangers and local community leaders. 4. Which factor best explained the likelihood that communities responded positively to rangers? Our final analysis aimed to identify the factors that best explained the likelihood that community members would respond positively to rangers. The number of positive responses per month was converted into binary data comprising months with a positive response (19962000 monthly totals; n = 81) and months without (1996-2000 monthly totals; n = 60) a positive response. This formed the dependent variable in a stepwise logistic regression analysis, using the forward stepwise procedure. The explanatory variables were significant factors identified from the univariate analyses. Areas of Bwindi were entered as categorical variables.

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

79

Results Most responses to rangers by community members were negative (Figure 4). Of these, most were complaints about crop raiding by wild animals (84.3%). There were fewer complaints about ICD benefits (10.0%) or about the national park (5.7%). Therefore, five years after the violent conflict that arose between local communities and park staff during national park gazettement and after MUP implementation, complaints about crop raiding by wild animals were the most common type of response that communities neighbouring Bwindi made to rangers. 60

Community response (%)

50

40

30

20

10

0 Very Negative

Negative

Neutral

Positive

Very Positive

Type of community response

Figure 4. Positive and negative responses by communities to rangers at Bwindi Impenetrable National Park from 1996 to 2000.

Area Most community responses to rangers in each area of Bwindi were negative (Pearson’s χ2 = 60.83; df = 12; p < 0.001; Table 2; NB: the west area was omitted from analysis because of low cell frequencies). Rangers in the north and centre areas of Bwindi received the highest proportion of negative responses. In comparison although rangers patrolling the east mainly received negative responses, the proportion of positive responses that they received was higher than other areas (Kruskal-Wallis χ2 = 18.25; df = 4; p < 0.01). The proportion of positive responses to rangers in the south and west areas was similar. The west area of Bwindi, where Mountain gorilla tourism was first established, was the only area where communities did not make very positive responses towards rangers, for example providing assistance with law enforcement activities. Therefore during the 1996-2000 period at Bwindi, rangers in the east received the highest number of positive responses by community members, whereas rangers in the north and centre received the highest number of negative responses.

Complimentary Contributor Copy

80

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Table 2. Type of community response to law enforcement rangers by area of Bwindi from 1996 to 2000 Response

Very negative Negative Neutral Positive Very positive

Area (%) North (n = 96) 14.6 64.6 7.3 7.3 6.3

Centre (n = 153) 21.6 60.8 7.8 3.9 5.9

East (n = 110) 4.6 40.0 12.7 21.8 20.9

South (n = 39) 20.5 46.2 10.3 7.7 15.4

West (n = 22) 9.1 54.6 27.3 9.1 0.0

Community Member Villages accounted for most of the responses made to rangers (89.3%). There were fewer responses by authorised resource users (5.7%) and local councillors (5.0%).

Villagers and Local Councillors Villagers made all of the very negative responses to rangers, except for one that was made by local councillors of village courts in the north who refused the trial of individuals arrested by rangers in the National Park. In general most responses by villagers and local councillors towards rangers were negative (Table 3). Authorised Resource User In contrast most responses by authorised resources users were positive. There were no occasions of very negative responses by authorised resource users where they refused to assist rangers with law enforcement activities or requested compensation because of crop raiding by wild animals (Fisher’s Exact Test = 69.84; p < 0.001). Table 3. Type of community response to law enforcement rangers by community member of Bwindi from 1996 to 2000 Response

Very negative Negative Neutral Positive Very positive

Community member (%) Villager Local authority (n = 374) (n = 21) 16.1 15.1 60.6 55.4 9.3 9.3 6.2 9.3 7.6 10.8

Resource user (n = 24) 0.0 12.5 12.5 37.5 37.5

Most responses by authorised resource users to rangers were made by beekeepers of the east (87.5%) and these beekeepers mainly responded positively towards rangers (66.7%). Authorised herbalists and basket makers of the north rarely interacted with rangers (8.4%) (Figure 5).

Complimentary Contributor Copy

81

Linking Protected Area Conservation with Poverty Alleviation in Uganda

None of the responses in the analysis were made by authorised beekeepers of the south, although these beekeepers did assist rangers with fire control. The 23 incidents when rangers received assistance by local communities with fire control were in 1999 (n = 13) and 2000 (n = 10). The assistance was given by villagers around the centre (n = 4), villagers and authorised beekeepers around the south (n = 8) and villagers and authorised beekeepers in the east (n = 11).

Community response (%)

100

80

60

40 Negative Neutral Positive

20

0 North

Centre

East

South

Area of Bwindi

Figure 5. Responses to law enforcement rangers by resource users of the harvest zone programme in areas of Bwindi from 1996 to 2000.

Therefore, in summary, most responses by villagers and local councillors to rangers were negative. In contrast, individuals directly benefitting from MUP, particularly beekeepers in the east, mainly responded positively towards rangers.

Year and Season There was no difference in the mean proportion of positive community response between years from 1996 and 2000 (Kruskal-Wallis χ2 = 4.02; df = 4; p > 0.05); months of the year (Kruskal-Wallis χ2 = 10.63; df = 11; p > 0.05); months of the rainy or dry seasons (z = -1.14; p > 0.05) or months of the farming season (Kruskal-Wallis χ2 = 1.24; df = 2; p > 0.05). Therefore year and season did not affect whether individuals responded positively or negatively towards rangers.

Patterns of Community Response The final model that best explained patterns of community responses to rangers comprised the three-way, significant interaction of response-type*area*community-member. The standardised lambda values revealed that community members in the north and centre of Bwindi were mainly associated with very negative and negative responses towards rangers

Complimentary Contributor Copy

82

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

and that, in contrast, community members in the east were mainly associated with very positive and positive responses towards rangers (Figure 6, Tables 4 and 5). Table 4a. Positive responses to law enforcement rangers by communities in areas of Bwindi from 1996 to 2000 Response

Area (%) North Centre East South West (n = 20) (n = 27) (n = 61) (n = 13) (n = 7) Very positive Report IA to outpost 0.0 0.0 5.1 5.5 0.0 Report IA to patrol 0.0 0.0 2.7 2.7 0.0 Assist investigation 6.3 5.9 7.7 12.7 0.0 Positive Positive comment 13.5 9.8 7.7 21.8 9.0 Appreciation 0.0 0.0 2.6 2.7 0.0 Neutral Report problems 1.0 2.0 5.1 10.0 22.7 Enquiry 0.0 0.0 2.4 0.0 4.5 Key: report IA to outpost (report illegal activities to rangers at the outpost); report IA to patrol (report illegal activities to rangers on patrol); assist investigation (assist rangers investigating illegal activities); positive comment (positive comments about the National Park and/or conservation); appreciation (appreciation for rangers assistance with problem animal control); enquiry (enquire about National Park issues).

Table 4b. Negative responses to law enforcement rangers by communities in areas of Bwindi from 1996 to 2000 Response

Area (%) North Centre East South West (n = 76) (n = 126) (n = 49) (n = 26) (n = 14) Very negative Refuse assist IAs 5.2 5.9 0.9 5.1 9.1 Refuse trial 1.0 0.0 0.0 2.3 0.0 Alert offenders 0.0 0.7 0.9 2.3 0.0 Request CR 8.3 15.0 2.7 10.3 0.0 Negative Complain CR 54.2 51.0 34.6 35.9 50.0 Complain NP 6.3 6.5 4.6 7.7 4.6 Complain ICDP 4.2 3.3 0.9 2.6 0.0 Key: refuse assist IAs (refuse to assist rangers investigating illegal activities); refuse trial (refuse to assist rangers with the trial of offenders); alert offenders (alert offenders inside the National Park to rangers on patrol); request CR (request compensation, vermin guards, or land purchase from National Park officials because of crop raiding); complain CR (complain to rangers about crop raiding by wild animals); complain NP (complain to rangers about living adjacent to the National Park); complain ICDP (complain to rangers about community benefit schemes of the ICDP)

Northern and Centre Areas: Complaints about Crop Raiding When villagers in the north and centre areas interacted with rangers, they mainly did so to complain about crop raiding by wild animals. Requests for assistance with crop raiding were fewer although higher around the centre than the north.

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

83

Proportion of positive responses

1.0

0.8

0.6

0.4

0.2

0.0 North (n=34)

Centre (n=46)

East (n=32)

South (n=17)

West (n=12)

Area of Bwindi

Figure 6. Mean+SE proportion of positive responses per month to rangers by communities in areas of Bwindi from 1996 to 2000.

Positive responses were few but included assistance with law enforcement: rangers did not receive reports of unauthorised resource use from community members, although communities did assist rangers when rangers asked for information on unauthorised resource use. In the north, most of this assistance was by villagers (83.3%) with some assistance by local councillors (16.7%). In the centre, most of the assistance was by villagers (55.6%) although there was a higher proportion of assistance by local councillors (33.3%) and some assistance by authorised resource users (herbalists and basket makers) (11.1%). Positive responses also included positive comments about the national Park. In the north villagers made most of the positive comments to rangers (66.7%) with fewer by authorised resource users (16.7%) and local councillors (16.7%). In contrast in the centre, local councillors (83.3%) made most of the positive comments with some by villagers (16.7%).

Eastern Areas: Assistance with Law Enforcement Most responses to rangers by communities in the east were positive. Unlike the north, centre and west, rangers in the east received reports of unauthorised resource use both on patrol and at the outpost when a community member travelled from their homestead to the outpost to report unauthorised resource use. Villagers (66.7%) made most of these reports with some by authorised beekeepers (33.3%). Both villagers (50.0%) and authorised beekeepers (42.9%) assisted rangers with investigations into unauthorised resource use. Both also made positive comments about the national park to rangers although most of these positive comments were made by authorised beekeepers (57.1%) with fewer by villagers (21.4%) and local councillors (21.4%). In comparison with other areas around Bwindi, in the east there were few incidents where community members refused to assist rangers with law enforcement. However both villagers

Complimentary Contributor Copy

84

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

(92.1%) and authorised beekeepers (7.9%) made complaints to rangers about crop raiding by wild animals.

Southern Areas: Complaints about Crop Raiding Similarly to northern and central areas, most interactions between communities and rangers were either villagers requesting assistance to control crop raiding or complaining about crop raiding. Of the few positive responses made, these comprised villagers reporting unauthorised resource use to rangers on patrol and at their outpost, and villagers (66.7%) and local councillors (33.3%) assisting rangers with their investigations into unauthorised resource use.

Western Areas: Refusals The only type of very negative responses to rangers in the west was villagers refusing to assist them with law enforcement. The west was the only area with no records of community members assisting rangers with law enforcement. The few positive responses to rangers by communities were villagers making positive comments about the national park. There were neutral responses whereby villagers (88.3%) and local councillors (11.7%) reported problems of the national park to rangers.

Discussion Our analyses showed that, five years after violent conflict between local communities and park staff when Bwindi was gazetted and after MUP, villagers complaining about crop raiding by wild animals was the most common type of interaction between rangers and local communities for the 1996-2000 period. Our analyses also showed that these complaints were particularly high in northern and central areas whereas, in contrast, villagers and authorised beekeepers in eastern areas responded more positively towards rangers. Year, month and season were not significant influences on whether communities responded positively or negatively to rangers. Our approach of examining daily interactions between rangers of a national park and the neighbouring communities therefore provides insight into local conflict issues and an indication of the factors that engender support for protected areas. For Bwindi we were able to assess causes of conflict and the influence of MUP in shaping local attitudes and behaviours towards park authorities.

Causes of Conflict Indicators of conflict were the negative and very negative responses that communities made to rangers. Very negative responses included incidents when community members

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

85

alerted offenders to an approaching patrol or refused to assist rangers with law enforcement. For example, in 1998, rangers recorded that villagers neighbouring central areas would not give information about hunters following an incident of bushmeat poaching in the national park. Also in 1998 around the west where villagers benefit from gorilla tourism, rangers found illegal pole cutting in the national park and noted “we rangers asked people who cut the poles but they refused to tell us”. Negative responses were community members complaining to rangers about an issue associated with the national park. From rangers’ descriptions, the complaints regarded three issues: crop-raiding by wild animals, ICD benefits and loss of forest resources such as, in 1996, the response to rangers patrolling eastern areas where beekeeping inside the park was allowed “the community were very annoyed and discontent, they were saying that since we had taken their bamboos and their firewood, now and then we should co-operate with them”. Our study showed that complaints about crop raiding dominated interactions between villagers and rangers over the five-year period of 1996-2000. We also showed that this type of conflict was most common in northern and central areas. Northern and central areas are known to experience high levels of crop raiding, particularly by baboons (Baker, 2004), and were associated with high levels of violent conflict during gazettement (Baker et al. 2011). These areas are still associated with conflict, for example in 2012, villagers of the centre constructed a road block to stop vehicles of conservation authorities and tourists passing through Bwindi to protest about the lack of benefits that they receive from the National park. Previous studies have demonstrated the negative influence of crop raiding by wild animals on relations between local communities and protected area authorities. This includes hostility between communities and authorities (Newmark et al., 1993; Hill, 1999; Nyhus, Tilson and Sumianto, 2000) and the undermining of efforts to gain local support for conservation (Infield, 1988; Archabald and Naughton-Treves, 2001). Indeed at Bwindi, crop raiding is known to adversely affect local attitudes towards the national park, particularly of poorer villagers (Blomley et al. 2010). Our study supports these findings, as villagers made a direct link between ICD benefits and livelihood impacts of crop raiding when, in 1996, villagers’ complaints to rangers included “people around the northern sector are not happy because the money of the Bwindi Trust is given to those who never had problems of the forest animals”. Furthermore, crop raiding affected the willingness of individuals to assist rangers with law enforcement. For example in 2000, rangers patrolling the centre recorded “we could not get any response on illegal activities, only people complaining about baboon damage”. However rangers did receive assistance with law enforcement from women and children who were guarding crops from wild animals. For example in 1998, rangers patrolling the centre recorded “a woman guarding from vermin told us that children were fishing inside the national park”, and, in 1996, rangers patrolling the east recorded “we were told by a young boy who was chasing monkeys from his garden that firewood collection always occurs on Sunday evenings”. This inconsistency reveals the complexities of community-park relations and highlights the need for conservation practitioners to understand this complexity and the different community groups involved. From rangers’ notes on the complaints of crop raiding that they received, crop raiding in itself was not the root cause of conflict. Rather, it was the perception by villagers that national park authorities had failed to reduce impacts on their livelihoods from crop raiding. This was despite benefits from ICD and efforts by UWA and ICD practitioners to reduce crop raiding around Bwindi.

Complimentary Contributor Copy

86

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Historical records document crop raiding around Bwindi since the early 1900s by gorillas and the assistance that authorities gave to local communities to reduce crop raiding (section x). There could therefore be an expectation that authorities should reduce crop raiding. Our findings suggest that directing ICD funds towards crop-raid control, particularly for northern and central areas, might reduce conflict and address conservation-poverty linkages by alleviating impacts on local livelihoods from the national park. However, there are many considerations. Firstly if the aims of crop-raid control interventions are to reduce conflict and alleviate livelihood impacts of crop raiding, it must be noted that areas where crop raiding is a major cause of conflict might differ from areas where villagers incur the greatest cost of crop raiding. Similarly, if crop-raid control interventions are to endanger local support for the national park and, in doing so, reduce unauthorised resource use, the association between individuals incurring the greatest cost of crop raiding and those undertaking unauthorised resource use must first be established. Finally, crop-raid control interventions should be implemented on good governance principles whereby villagers are engaged with the decisionmaking process and have ownership of the type of intervention, particularly for the interventions to reduce conflict in the long-term. Therefore, ICD efforts to reduce crop raiding could resolve a major cause of conflict at Bwindi and contribute towards poverty alleviation. However, the complexities surrounding this issue and conservation-poverty linkages must first be fully identified and understood.

Local Community Support for the National Park Very positive and positive responses by community members to law enforcement rangers indicated local support for the national park. Very positive responses included community members reporting unauthorised resource use to rangers and assisting rangers to investigate unauthorised resource use. For example, in 1998, authorised beekeepers of the east reported snares in their harvest zone to rangers, and the rangers recorded “the beekeepers were not happy with this activity, which is carried out in their zone. They gave us two porters of their society to lead us to those snares. All snares we found were new and we talked with these porters to organise another patrol so they can lead us to other suspected places in the same area.” Positive responses included positive comments about the national park, such as the response by an authorised beekeeper to rangers patrolling the east in 1996 “one man who was also a beekeeper member told us that people are ready to look after the park as they promised themselves as beekeepers, we thanked the beekeepers bordering the area and encouraged them to continue with the same spirit”. During the period of gazettement from 1989 to 1992, villagers in eastern areas of Bwindi undertook a series of violent attacks on rangers following the arrest of fellow villagers for transporting cattle through Bwindi to sell in Rwanda (Baker et al. 2011). Our results showed that, from 1996 to 2000, authorised beekeepers and villagers of eastern areas accounted for the highest proportion of all positive and very positive responses that rangers recorded, and that their support was largely to assist rangers with law enforcement. This finding indicates a positive impact of MUP on the conservation attitudes and behaviours of both individuals directly benefitting from MUP and individuals not directly benefitting from MUP but living in a MUP village. The role of authorised resource users in protecting Bwindi and reporting unauthorised resource use was emphasised during implementation of MUP (Bensted-Smith et

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

87

al., 1995; Wild and Mutebi, 1996). Therefore the assistance by beekeepers to rangers could be expected although individuals not directly benefitting from MUP also responded positively to rangers. This could have resulted from local benefits that MUP generated including production of honey and the continuation of a traditional practice. However, authorised resource users and villagers of other MUPs did not respond as positively towards rangers as those in eastern areas. This difference in conservation support between different MUP areas could reflect the difference between these areas in MUP implementation. Eastern beekeepers were the first communities neighbouring Bwindi to be granted access to the forest, which was the year following gazettement (Bensted-Smith et al., 1995; Wild and Mutebi, 1996). MUP implementation in other areas began three years after gazettement and the process was delayed by organisational failure (Bensted-Smith et al., 1995), which created frustration amongst villagers with conservation authorities (Blomely, 2003). Therefore, the relatively quick implementation process for eastern beekeepers and the re-establishment of forest ownership shortly after gazettement appear important for securing local support for the national park.

Daily Interactions between Local Communities and Rangers as Indicators of Conflict and Support Monitoring projects in multi-disciplinary terms of ecological and socio-economic impacts is important to determine whether the ICD approach can protect wildlife (Larson and Svendsen, 1996) and alleviate poverty. However, choosing socio-economic indicators is difficult because of the complexities involved (Kleiman et al., 2000). Attitudes are often used to determine local support for conservation and thus social impacts of conservation policy (Straede and Helles, 2000; Mehta and Heinen, 2001; Zhang and Wang, 2003), and factors that influence community attitudes towards protected areas are well documented, including wealth and education (Fiallo and Jacobson, 1995; Sah and Heinen, 2001; Holmes, 2003) and crop raiding by wild animals (Hill, 1998; Mehta and Heinen, 2001). However, positive community attitudes towards protected areas do not necessarily translate into conservation benefits (Badola, 1998; Straede and Helles, 2000; Infield and Namara, 2001), which raises the question as to the efficacy of attitudes as indicators. Furthermore, evaluating conservation policy requires an understanding of behavioural change particularly regarding resource use (Holmes, 2003), which limits the use of attitude surveys to evaluate conservation policy. We used daily interactions between community members and rangers over a five year period to indicate conflict and local support for the national park and enable comparisons of positive and negative responses towards rangers between individuals directly benefiting and not directly benefiting from MUP. There were factors outside the scope of our analysis that possibly influenced how communities responded to rangers. These included law enforcement: several arrests in one area might have fuelled resentment amongst villagers that lead to their refusal to assist rangers with law enforcement activities; ranger patrol effort likely differed between areas during the study period; and, the extent that rangers asked villagers for assistance with law enforcement might have differed between areas. Also the amount and type of ICD interventions differed between areas, as well as political pressures, markets, land use, population density and other external factors. A further limitation is that the data comprised descriptive reports by law enforcement rangers. We verified the data by fieldwork

Complimentary Contributor Copy

88

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

and incorporated the local community perspective from focus group discussions. However, the data primarily reflects rangers’ perceptions of their interactions with community members. Whilst acknowledging these limitations, our study does show that crop-raiding dominated ranger-community interactions and villagers who received the first MUP shortly after gazettement demonstrated the highest level of support for the national park. Our study also shows the use of historical records for social assessments within conservation management and that developing typologies of park-community interactions gives insight into the causes of conflict and factors that engender local communities to support protected areas.

8. ASSESSING GOVERNANCE OF LOCAL COMMUNITY PROJECTS BY THE BWINDI MGAHINGA CONSERVATION TRUST A Conservation Trust Fund for Bwindi and Mgahinga National Parks Table 5. Funding categories of BMCT Category Park Management Conservation / ecological research Communities

Activities that receive funding Investment in UWA projects, training or equipment Funding to research projects and ecological monitoring by the Institute of Tropical Forest Conservation (ITFC) or affiliated researchers Common goods (communal) projects Projects that benefit communities as a whole, for example provision of classrooms, health units, laboratories and dormitories, community water harvesting tanks and gravity water schemes Conservation projects with communities Projects that aim to conserve natural resources through providing communities with alternative resources, for example tree planting and energy saving stoves Projects that reduce conservation costs incurred by communities from crop-raiding of wild animals, for example planting the Mauritius live fence for problem animal management Income/livelihood projects Projects that aim to improve household income, for example support for potato growing, mushroom growing, livestock, beekeeping outside of the national park, fish farming, handicrafts for sale, vegetable growing, Village Saving and Loan Associations Batwa support Projects targeted to improve household livelihoods and well-being of the Batwa including land purchase and settlement, income generating activities and education

Bwindi Mgahinga Conservation Trust (BMCT, formerly MBIFCT) was created in 1994 by a GEF World Bank project to support biodiversity conservation of Bwindi and Mgahinga Gorilla National Park (MGNP). The establishment of BMCT was to overcome the barrier of the need for long-term funding and support for ICD success. Therefore with its (currently) long-term funding commitment to ICD interventions, BMCT is a major ICD practitioner for both mountain gorilla national parks in Uganda. The Trust mandate is to directly support national park conservation by funding research and park management activities and to support conservation through local community development programmes. Project activities of BMCT began in 1997 and continue to the

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

89

present day. In the 15 years since BMCT began its activities, it has funded various projects under three broad categories of park management, conservation / ecological research and communities (Table 5). Community development programmes receive most of BMCT’s funds. Allocation of these funds is in accordance with the following goal: Provide alternative means of meeting needs which were traditionally met by harvesting forest resources (e.g. timber, poles, meat, cash income). Among the types of activities likely to be funded are beekeeping (including marketing of products), agroforestry and woodlots, dairy and poultry production1, and ecotourism services and facilities. These economic activities will help compensate the communities for the loss of their traditional access to these resources when the forests were gazetted as protected areas. They will also help discourage illegal exploitation and burning of the forests, which the authorities cannot completely eliminate through simple enforcement, by providing alternatives and by fostering a positive attitude among the communities toward these national parks as a source of concrete benefits. The cooperative approach to managing the funds will also provide an opportunity for the different stakeholders to work together to identify and achieve common objectives.…The Trust is meant to provide incremental support, complementing but not displacing funds from the GOU and other donors…The project also represents an experiment in application of two important principles for biodiversity conservation: (I) including community representatives as full partners in decision-making, as a means of gaining community support for conservation and "ownership" of the project, and (2) the use of a Trust Fund as a mechanism for providing reliable, long-term funding for conservation activities. (GEF 1995)

In addition, community projects that receive BMCT funding must be projects that: (i) are proposed by established local community groups; (ii) have a demonstrable positive impact on conservation of the parks and their biodiversity (e.g. non-consumptive utilization of forests such as eco-tourism; development of substitutes for vulnerable resources); (iii) are consistent with UNP [now UWA] policies and park management plans; (iv) meet agreed criteria of social and environmental soundness, equitability and transparency; (v) include a matching contribution in cash or kind by the proposing group; and (vi) include provisions and arrangements for accountability and long-term sustainability (ibid).

These criteria were developed to represent aspects important to ICD success: alternative livelihood provision, compensation for lost resources, positive attitude development, sustainability, collaboration, ownership and governance. However to date there has been no evaluation of whether BMCT has achieved these aspects, particularly the principles of good governance within its operations and funded projects. Furthermore, while general evaluations of ICD at Bwindi indicate that projects developed and undertaken by BMCT have had a positive impact on the attitudes of local communities, all identified the need for a detailed assessment of BMCT impacts on national park conservation and rural development goals 1

Since this was written, one of the Trust’s core activities using endowment funds is community agriculture especially provision of improved seeds and extension work

Complimentary Contributor Copy

90

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

(Craig and Malpas, 2007; World Bank, 2007; Blomley et al. 2010). This lack of monitoring and evaluation limits the ability of BMCT, as an ICD practitioner, to gauge its success in conservation and poverty alleviation goals, to understand whether it has achieved good governance and, in doing these, make a meaningful contribution to the ICD debate.

Governance Assessment After 15 years of projects, BMCT elected to conduct an impact assessment of its activities with park management, research and communities. It also decided that the starting point for the impact assessment would be governance. Therefore the aim of this study was to assess whether BMCT had embedded the principles of good governance into its operations and projects, particularly to determine whether community representatives were full and genuine partners in the decision-making process and felt a sense of ownership of BMCT projects.

Methods Attitude surveys were designed for stakeholders in BMCT community projects that included local government officials, local leaders and villagers. The questions were first to establish livelihood profiles, which for the Batwa included land, housing and education, and then to explore attitudes on governance issues of BMCT’s activities (both common goods and livelihood projects). Interviews were held with local government officials at parish and subcounty levels. Attitude surveys of villagers and local leaders neighbouring the national parks were conducted in two villages per parish: one with a BMCT project and one selected at random with the criteria that the village was not neighbouring the first village and, where possible, one of the two villages bordered a national park. There were ten informants per parish that included (as available) one local leader, one head school teacher, two Trust project participants and five villagers selected at random.

Initial Findings At the time of writing, the study was in the final stages. Our initial findings on governance issues are presented here in two sections: project governance, and grassroots representation (the ability of local people to be involved in the initial stages (application, project selection) of Trust interventions).

Project Governance Two findings have been immediately evident from the attitude surveys. First that most respondents rated their involvement in BMCT projects as being ‘very important’ (Figure 8.1). Second that most respondents who benefitted from BMCT projects did feel involved (figure 8.2). Recipients of BMCT livelihood projects felt involved in both the project design and implementation. Of note, in order to compare the Trusts’ approach to other organizations and

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

91

agencies, we had a particular focus on water projects (categorized under common goods projects) given the Trust had just completed work on a large gravity scheme. Results suggested recipients of BMCT water projects felt more involved in project governance than recipients of other water projects (Figures 8.3-8.6).

Village attitudes on the importance of involvement 200

161

150 100 50 2

2

8

Don't know

Not really

Somewhat

0 Very

Figure 7. Attitude from respondents on local involvement in common goods projects. Villagers want to participate in decision-making processes that relate to projects impacting them (n=173)

Sense of Trust communal project involvement 120 100 80 60 40 20 0 No

Yes

Figure 8. In BMCT projects, a vast majority of respondents felt they were involved in the project design and implementation (n=129).

Complimentary Contributor Copy

92

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Should villagers be involved if they are not 80 70 60 50 40 30 20 10 0 Don't Know

No

Yes

Figure 9. Project involvement is very important to local people, highlighting the need for all projects to ensure the opportunity to participate. (n=72).

Sense of involvement by recipients of water schemes 45 40 35 30 25

Yes

20

No

15 10 5 0 Baynara Scheme

All water projects

Figure 10. Using a specific type of common goods project, villagers participating in the Trust's Banyara Gravity Water scheme felt involved in project design, whereas more generally water projects are not as inclusive. (n=83).

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

93

Figure 11. Participants in Trust livelihood projects nearly unanimously felt involved in the projects. (n=50).

Did project participants feel able to speak their minds?

Yes No Don't Know

Figure 12. Similarly, livelihood project participants were comfortable to engage Trust staff openly during project implementation. (n=48).

Interviews with villagers and local government officials also showed that there is some level of ownership in common good and livelihood projects. Local government officials in particular (20 of 23 respondents) felt able to participate with BMCT community officers and the Local Community Steering Committee (LCSCs) and involved in BMCT projects, particularly to voice opinions on which projects should be chosen. One of those who did not feel so involve responded “BMCT invite us for meetings but we are not involved in project implementation and monitoring”. This theme of lack of monitoring and evaluation was evident from the beginning of the survey, as BMCT itself noted that they wanted to use this study to help develop an M/E program. What was interesting is that it was a theme that came up repeatedly throughout the surveys in the villages and with government officials. The

Complimentary Contributor Copy

94

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

quoted government official said to ensure local government involvement, the Trust should “involve me right from planning project selection, implementation and monitoring for project sustainability”. Various suggestions for monitoring include providing money for LCSC to monitor projects, involving local government to monitor projects (leaders felt if villagers knew they were monitored they may work harder), allocating more money in the Trust budget for BMCT community officers to monitor projects, and finally develop a joint monitoring and evaluation team that includes BMCT, local government, and villagers in monitoring projects and activities. Therefore this study provides evidence that BMCT’s policies on governance issues of collaboration and ownership with local stakeholders have been incorporated into its project activities. However two aspects of governance within BMCT operations need consideration. Firstly BMCT must ensure that its operations with local government officials are carefully conducted to make sure that its good relationship with government does not result in its projects providing special benefits for local government officials and their supporters, or indeed that local politics begins to interfere with normal governance processes of BMCT projects. Political bias in the misuse of conservation funds has been documented (Blomley et al. 2010) around Bwindi in other projects. The Trust has successfully avoided this issue within its 15 years of operation and, to build on this good foundation, should establish and implement measures to mitigate this risk and be transparent in its efforts to do so.

Grassroots Representation Our third major finding within this governance study came from the structure set up by the Trust to ensure community participation in the Trust. The Trust works with communities through the LCSC (Local Community Steering Committee), a structure that provides for 5 community representatives (one from each district, a woman’s representative, and a Batwa representative) on the Trust’s Board. The 5 Board representatives are elected by a system of LCSCs which contain representatives of each of the 15 sub-counties across BMCA. These sub-county LCSCs are elected every 3 years by a panel of temporary parish-level LCSC representatives that are disbanded after the sub-county elections. The LCSCs communicate with local government, but are a separate entity, which provides the Trust with a mechanism with at some separation from local political influences and costs. The LCSC system is used for both communal goods and livelihood projects. The Trust engages the LCSC representatives to publicize livelihood grants both through local government chains and public radio messages, and uses the LCSC as the mechanism to communicate with the people around BMCA. Community members (often in the form of a group) interested in developing a project then write proposals and those that pass the scrutiny of LCSC are funded. Where possible the LCSC member provides assistance to groups needing capacity support, and in some cases encourages stronger individuals within the community to partner with poorer individuals to strengthen applications. Although this system was designed to provide local representation and encourage local participation (while simultaneously being financially feasible to operate), the study revealed the system has two major shortcomings: lack of local participation/representation at the village level, and a lack of local LCSC oversight. LCSC representatives are not paid and are not financially supported to be mobile. Because of this the LCSC representatives often remain at the sub-county and do not work at the village level where most projects occur. The result, for many of the people that were interviewed, is that most villagers did not know about the

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

95

LCSC representative and, consequently, did not feel represented (Figures 8.7-8.11). 56 of the 108 informants who wanted to participate in Trust livelihood projects was that they did not know how to apply for them and thus did not have the opportunity to become involved. This is a significant barrier that could prevent BMCT from targeting livelihood projects at poorer members of local communities, especially those with weak links to local government. In addition, as most LCSC residences are not located near the park, those villages bordering the park (who suffer the highest costs of conservation) are least likely to be reached in communications/interactions with the LCSCs. The second shortcoming noted by informants is the lack of term limits and a local oversight system for the LCSC representatives. Several LCSC members have served since the Trust was created. Although having LCSC members who are well informed of conservation goals, understand the Trust, and show willingness to serve, there are consequences of having entrenched representatives: stagnation, lack of local accountability, and in some cases, project recipient bias away from those who should be targeted in ICD projects. More importantly, the lack of a system of local transparency perceived by villagers can impact how the Trust and its activities are viewed. One local government representative reported that villagers think he is taking money from the Trust as ‘only one out of 100 projects are funded’ due to the lack of Trust funds available for livelihood projects. If the system was more transparent these challenges may not exist. The consequences of these two LCSC system weaknesses are that good governance is not achieved at the village level during the project identification stage and that BMCT do not reach the poor and marginalised. This barrier that approximately half of our interested informants faced however is an implementation issue in selected parishes, rather than the failure of BMCT livelihood projects to positively impact on household income, and should be addressed accordingly. Despite the governance challenges identified, BMCT activities were well received by communities, local government and park management. All stakeholders stated a desire to see more activities from BMCT, demonstrating the positive impact that it has had on the region.

Figure 13. Almost half of all villagers interviewed in the area of Trust operations did not know the LCSC system that provides representation and awareness of Trust projects.

Complimentary Contributor Copy

96

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Do the LCSCs serve as a tool for representation and ownership in Trust activities? Don't Know No Yes

Figure 14. The LCSC system provides some ownership, however the study suggests that there is the need to improve this governance system. (n=114).

Does the LCSC system need strengthening? Don't Know

No

Yes, 95%

Figure 15. When asked directly, nearly all informants said the Trust needs to improve the LCSC system. Further probing revealed that people felt the representation needed to move from the subcounty level to the village level. (n=180).

How to strengthen the LCSC system 40 30 20 10 0

Figure 16. Most respondents felt devolving representation to the village level, oversight, and more LCSC governance awareness were the best ways to improve the LCSC system. (n=67).

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

97

How to strengthen LCSC according to Local Government 10 8 6 4 2 0

Figure 17. From local government's perspective, providing finances to existing LCSC members and devolving LCSC governance to the village level would improve the LCSC system. (n=24).

Applying Lessons Learned: BMCT Management Response Governance challenges and policy suggestions are much easier to identify and recommend than to actually implement. The Trust has, overall, done a good job in developing a system of representation that avoids government influence and is financially balanced with providing communities with funds to implement projects. Their response to the findings of the study of the LCSC system reflects this: “BMCT uses the existing local government structure and even the social sector like churches to publicise the projects. During project selection the civil society as well as the local government elected officials at parish level and the government extension staff are involved in project identification appraisals and final selection with the approval by the LCSC for projects... I believe the rigorous process this involves can be improved but everything has been done to make transparent and inclusive of all the target communities (Mwine, pers comms).” Although the study found areas for improvement, the Trust is conscious of the need for good governance and the need to apply adaptive management when weaknesses are found. Two of the main challenges with projects in general addressing governance issues is that it is time consuming and very costly. Is it feasible to set up an expensive system of representation when only one livelihood project per 5000 people is allocated? For the Trust, the two challenges that the study has highlighted to be addressed include local representation and local oversight. Below outlines the suggestions to BMCT to address these governance challenges. 1. Renovate the LCSC system. Given funding constraints, developing an LCSC system down to the village level seems unlikely. Thus, the Trust needs to: a. Improve LCSC transparency and monitor activities

Complimentary Contributor Copy

98

J. Baker, R. Bitariho, A. Gordon-Maclean et al. b. Impose term limits on LCSC c. Facilitate LCSCs to hold regular meetings with LC1s and stretcher groups 2. Develop a governance component for the M/E system. Ensuring the Trust practices good governance is key for long-term sustainability of projects. Although the Trust does engage in good governance, there are cases that came up during our interviews that suggest a system to monitor projects and staff would help ensure this. Consider the following: a. Are key stakeholders involved from decision-making to project design to implementation to monitoring? b. Are benefits equitable and free of corruption/elite capture? c. Attending meetings does not equate to meaningful participation d. Make transparent the system for choosing and awarding livelihood projects e. Have beneficiaries been involved in sourcing materials?

In part because funding constraints do not allow for grassroots LCSC representatives at the community level, we are also recommending that the Trust scale back their livelihood project activities and focus on projects with broader impact that can include communities in decision-making. Those that still impact household livelihoods might include trade schools, savings and loans trainings, and agricultural/animal husbandry trainings.

Conclusion The principles of good governance are the foundation for ICD and must be embedded within the operations and activities of ICD practitioners. However monitoring and evaluation efforts tend to focus on conservation and rural development impacts of ICD, omitting governance. This typically involves evaluating whether the ICD approach has reduced unauthorised resource use or improved the socio-economic wellbeing of local beneficiaries. The lack of monitoring on governance issues limits our understanding of successes and failures of ICD with the result often being recommendations that ICD projects need to be better targeted towards the poor and marginalised. While improving ICD targeting is important, ensuring that local stakeholders are engaged in the decision-making process and feel a sense of ownership of projects is vital for the long-term success of ICD. For ICD practitioners, establishing a strategic direction on governance is the first step towards embedding good governance within operational activities and projects. As a major ICD practitioner of Uganda’s mountain gorilla national parks, BMCT has established good governance principles within its policies and now, after 15 years of operations, seeks to determine whether these principles have been achieved. As well as providing an essential knowledge-base from which to assess BMCT impacts on national park conservation and poverty, assessing governance aspects of BMCT projects will indicate how well BMCT is performing as an ICD practitioner. This study showed that most local stakeholders perceived that they are involved in BMCT projects, which demonstrates that BMCT has achieved a good level of stakeholder participation and local ownership of BMCT projects. The study also showed that stakeholders placed a high value on their participation in BMCT projects, which indicates the importance of governance issues to BMCT project success. However the study identified a barrier

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

99

preventing villagers becoming involved with and benefiting from BMCT activities. Therefore BMCT delivers its principles of good governance when projects are funded although, before that stage when villagers are to propose projects for funding, the LCSC system needs improving to ensure that opportunities to benefit from BMCT projects are equal and that BMCT can have greater impacts on rural development and national park conservation by reaching the poorer villagers neighbouring the national parks. Monitoring and evaluating governance can be difficult, particularly to understand power relations between stakeholders and how powers are exercised. This study employed attitude surveys to determine the perceptions of respondents on two aspects of governance – involvement with decision-making and project ownership. While attitude surveys are a starting point to more complex social research, this survey identified a key barrier to BMCT projects that can now be addressed and, in doing so, demonstrates the importance of assessing stakeholder perception on governance. Nonetheless, to fully understand governance within the context of ICD and national park conservation, there is a need to develop simple and robust measures for ICD practitioners to examine governance issues related to their operational performance and project implementation.

9. IMPROVING POLICY AND PRACTICE OF LINKING NATIONAL PARK CONSERVATION WITH POVERTY ALLEVIATION Linkages between national park conservation and poverty alleviation are positive and negative. Positive linkages are the contribution that conservation efforts make towards poverty alleviation and the benefits that poverty alleviation activities generate for national parks. Negative linkages occur when the establishment and maintenance of national parks exacerbate poverty and when economic development activities contribute towards biodiversity loss (Roe and Elliott, 2005).

Establishing a Conservation-Poverty Alleviation Policy Framework A starting point for achieving conservation though poverty alleviation is the policy change needed for conservation policy to take better account of poverty concerns, for development policy to take better account of biodiversity concerns and for both to pay attention to human rights. In this chapter, we described developments in Uganda’s conservation policy framework. Uganda has established a framework for achieving convergence between conservation and poverty alleviation objectives, and has complemented its policies with laws and institutions to achieve the convergence. Furthermore, its policies define the government’s aim to redress the inequitable distribution of costs and benefits of national park conservation and facilitate interventions to simultaneously pursue conservation and poverty alleviation objectives.

Complimentary Contributor Copy

100

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

The Historic Context of Natural Resource Management We assessed how the historical context of national resource management influences issues currently faced by national park managers. In Uganda these included community norms and individual beliefs regarding natural resource use, local expectations of assistance by authorities to reduce crop raiding by wild animals and local perceptions of ownership of the national park. Therefore, understanding historic relations between communities and authorities during different periods of natural resource governance, and how local people perceive their livelihood links to natural resources, will facilitate dialogue and learning on conservation-poverty linkage in order to promote best practice to policy-makers and practitioners.

Livelihoods and Wellbeing Another consideration is that, while policies for national park conservation are changing to better articulate pro-poor approaches, the management of national parks remains founded on law enforcement regimes to reduce illegal activities. Overcoming this legacy of historic natural resource management by moving away from concepts of ‘crime’ and ‘illegality’ is vital for national parks such as Bwindi where the aim is to align conservation efforts with poverty alleviation activities. Our approach is to define unauthorized resource use as an indicator of the different needs and uses of a national park by people, and of the governance challenges and limitations of national park management to balance people's needs with biodiversity conservation.

Evaluating Governance Finally, surveys of local attitudes towards national park conservation are useful to help gauge success of conservation efforts. Evaluating governance however, is vital to understand ICD success because whether poor people conserve or over-exploit biodiversity is dependent on specific circumstances and contexts – particularly the influence of governance (Roe and Elliott, 2005). Yet despite recognition of the need to engage local people in national park conservation, evaluations of whether good governance has been achieved are rare. This leaves conservation authorities with no benchmark to measure the governance aspects of their work, particularly on communications with local communities and the power relationships that dictate decision-making.

Summary Therefore, there are many challenges to achieving conservation through poverty alleviation. ICD is an approach for national park management to link conservation action with poverty alleviation, particularly by addressing local people's needs while reducing the resource use behaviours that threaten conservation. In this chapter we described identified key

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

101

factors to overcome the challenges and illustrated these with case studies of Uganda and Bwindi. For efforts such as those in Uganda to continue improving the policy and practice of conservation through poverty alleviation, promoting dialogue and fostering learning on conservation-poverty linkages is more important than ever.

REFERENCES Abbot, J.I.O and I. Guijt (1998) Changing views on change: participatory approaches to monitoring the environment. IIED SARL Discussion Paper No.2, IIED, UK. Adams, W.M & Hutton, J., 2007. People, Parks and Poverty: Political Ecology and Biodiversity Conservation. Conservation Society 5:147-183 Agrawal, A & Redford, K., 2009. Conservation and Displacement: An Overview, Conservation Society 7 (1):1-10. Alpert, P. (1995) Applying ecological research at integrated conservation and development projects, Ecological Applications, 5, 4: 857-860. Arnold, J.E.M, & Perez, R.M., 2001. Can non-timber forest products match tropical forest conservation and development objectives (analysis), Ecological economics 39; 437- 447. Archabald, K., and L. Naughton-Treves (2001) Tourism revenue-sharing around national parks in Western Uganda: early efforts to identify and reward local communities, Environmental Conservation, 28, 2: 135-149. Babulo, B, Muys, B, Nega, F, Tollens, E, Nyssen, J, Deckers, J, Mathijs, E, 2008. Household Livelihood strategies and forest dependence in the highlands of Tigray, Northern Ethiopia, Agricultural Systems 98: 147-155. Baker, J. Milner-Gulland E.J. & Leader-Williams, N. (2012) Park gazettement and integrated conservation and development as factors in community conflict at Bwindi Impenetrable Forest, Uganda Bensted-Smith, R., Infield, M., Otekat, J. and Thompson-Handler, N., 1995. Review ofthe multiple-use (resource sharing) programme in Bwindi Impenetrable NationalPark. A report to CARE-Development Through Conservation Project, Kabale. Bitariho R, Mugyerwa B, Barigyira R and Kagoda E., 2004. Local people’s attitudes and new demands since inception of multiple use programme in Bwindi Impenetrable National Park, S. W Uganda, Unpublished report, Institute of Tropical Forest Conservation, Kabale. Bitariho, R., Kagoda, E., Barigyira R.,Amanya, S., Safari., C., 2006.The potential supply of forest resources demanded by Batwa from Bwindi Impenetrable National Park, S.W Uganda, Unpublished report, Institute of Tropical Forest Conservation, Kabale. Bitariho, R & Barigyira R, 2009. The potential supply of plant resources for local community use. in Bwindi Impenetrable National Park Beekeeping Zones, S. W Uganda. A consultancy report to CARE-EEEGL,CARE INTERNATIONAL, Kachiru, Kigali, Rwanda. Blomley, T., Namara, A., McNeilage, A., Franks, P., Rainer, H., Donaldson, A., Malpas, R., Olupot, W., Baker, J., Sandbrook, C., Bitariho, R. and Infield, M., 2010 Development and gorillas? Assessing fifteen years of integrated conservation and development in south-western Uganda, Natural Resource Issues No. 23. IIED, London.

Complimentary Contributor Copy

102

J. Baker, R. Bitariho, A. Gordon-Maclean et al.

Borrini-Feyerabend, G. (ed) (1997) Beyond fences: seeking social sustainability in conservation. IUCN, Gland, Switzerland. Boot, R.G.A. and R.E. Gullison (1995) Approaches to developing sustainable extraction systems for tropical forest products, Ecological Applications, 5, 4: 896-903. Butynski T.M., 1984. Ecological survey of Impenetrable (Bwindi) Forest, Uganda, and recommendations for its conservation and management. New York Zoological Society Report, New York. CBD., 1993. Convention on Biological Diversity, www.cbd.int/convention Cunningham, A.B., 1996. People, Park and Plant Use, Recommendations for multiple use zones and development alternatives around Bwindi Impenetrable National Park, Uganda. A report for people and plants initiative, division of ecological sciences, UNESCO, 7 place de fontenoy, Paris. Cunningham A. B., 2001. Applied ethnobotany, people, wild plant use and conservation, WWF, Earthscan Publications Ltd, London. Davey, C., Peters, C.M, and Byarugaba, D., 2001. Participatory review of the Multiple Use Programme, Bwindi Impenetrable National Park and Mgahinga Gorilla National Park, CARE-Uganda and Uganda Wildlife Authority, Kampala. FAO., 2006. Better forestry, less poverty, a practitioner’s guide. Food and Agriculture Organization of the United Nations,FAO forestry paper 149, Rome. Freese, C.H., 1997. Harvesting Wild Species, Implications for Biodiversity Conservation. The Johns Hopkins University Press, 2715 North Charles street, Baltimore, Maryland. Ghazoul, J & Sheil, D, (2010). Tropical rain forest ecology, diversity, and conservation. Oxford University Press, Great Clarendon Street, Oxford. ITFC., 1999. The potential supply of weaving and medicinal plant resources in the Proposed kifunjo/masya multiple-use zone of Bwindi Impenetrable National Park, S.W Uganda. Unpublished report, Institute of Tropical Forest Conservation, Kabale. Kaimowitz, D &Sheil, D., 2007.Conserving What and for Whom? Why Conservation Should Help Meet Basic Human Needs in the Tropics,Biotropica 39(5): 567–574. Kremen, C., I. Raymond and K. Lace (1998) An interdisciplinary tool for monitoring conservation impacts in Madagascar, Conservation Biology, 12, 3: 549-563. Multiple use MoU, 1994. Memorundum of Understanding Between Uganda National ParkBwindi Impenetrable National Park and the People of Rutugunda Parish, Kanungu District. An agreement concerning collaborative forest use and management at Bwindi Impenetrable National Park, Kanungu. Namara, A., 2006. From Paternalism to real partnership with local communities? Experiences from Bwindi Impenetrable National Park (Uganda). Africa Development 31 (2):39-68. Ndangalasi, H, Bitariho R, Dovie D. 2007. Harvesting of non-timber forest products and implications for conservation in two montane forests of East Africa,Biological Conservation, 134 (2), 242-250. Newton, A.C., 2007.Conservation of Tree Species through Sustainable Use: How Can it be Achieved in Practice? Oryx, 42 (2):195-205. Larson, P., M. Freudenberger and B. Wyckoff-Baird (1997) Lessons learnt from the field: a review of World Wildlife Fund’s experience with integrated conservation and development projects 1985-1996. World Wildlife Fund, Washington D.C., USA.

Complimentary Contributor Copy

Linking Protected Area Conservation with Poverty Alleviation in Uganda

103

Peters, C. M, 1994. Sustainable Harvest of Non - Timber Plant Resources in Tropical Moist Forest:An Ecological Primer. USAID Biodiversity Support Programme, Washington D.C. Plumptre, A.J., M. Masozera, P.J. Fashing, A. McNeilage, C. Ewango, B.A. Kaplin, and I. Liengola (2002) Biodiversity Surveys of the Nyungwe Forest Reserve in S.W. Rwanda. Wildlife Conservation Society Working Papers No. 18, WCS, USA. Sandbrok, C.G, 2010. Local economic impact of different forms of nature-based tourism. Conservation Letters, 3 (1):21–28. Scott, P., 1998. From Conflict to Collaboration, People and Forests at Mount Elgon, Uganda. IUCN-Eastern Africa Regional Office, Nairobi Ticktin, T., 2004. The ecological implications of harvesting non-timber forest products. Journal of Applied Ecology, 41: 11-21. UWA, 2001. Bwindi Impenetrable National Park, Mgahinga Gorilla National Park (Bwindi/Mgahinga Conservation Area), General Management Plan (July 2001 – June 2011), Uganda Wildlife Authority, Kampala, Uganda. UWA, 2012. Bwindi Impenetrable National Park, Mgahinga Gorilla National Park (Bwindi/Mgahinga Conservation Area), Draft General Management Plan (July 2012 – June 2022), Uganda Wildlife Authority, Kampala, Uganda. Wells, M.P., K.E. Brandon and L. Hannah (1992) People and parks: linking protected area management with local communities. The World Bank, WWF and USAID, Washington, DC. Wells, M.P. and K.E. Brandon (1993) The principles and practice of buffer zones and local participation in biodiversity conservation. Ambio, 22, 2-3: 157-162. Wild, R, & Mutebi, J., 1996. Conservation Through Community Use of Plant Resources: Establishing Collaborative Management at Bwindi Impenetrable and Mgahinga Gorilla National Parks. People and Plants Working Paper 5. UNESCO, Paris. Wild, R., 2001. Plants from the Park: Establishing community harvesting of plants as a conservation tool at Bwindi Impenetrable and Mgahinga National Parks, Uganda Mphil Thesis, Dept of Botany, University of Capetown. World Bank.,2001. A revised forest strategy for the World Bank Group. Draft. Washington, DC. Woodroffe, R., Thirgood, S., & Rabinowitz (eds)., 2005. People and Wildlife: Conflict orCoexistence? Cambridge University Press, Cambridge. Wunder, S. (2000) Ecotourism and economic incentives - an empirical approach. Ecological Economics. 32, 3: 465-479. Vedeld, P., .Angelsen, A.,Sjaastad, E., &Kobugabe, G.B., G., 2004. Counting on The Environment, Forest Incomes and the Rural Poor. Environment Economics Series paper 98. World Bank. Washington DC.

Complimentary Contributor Copy

Complimentary Contributor Copy

In: National Parks Editor: Johnson B. Smith

ISBN: 978-1-62618-934-8 © 2013 Nova Science Publishers, Inc.

Chapter 3

COMMUNITY-BASED CONSERVATION: AN INSTITUTIONAL APPROACH TO ASSESSING BIODIVERSITY CONSERVATION EFFORTS AT BARDIA NATIONAL PARK IN NEPAL Shova Thapa Karki Lecturer Business and Management Jubilee Building, University of Sussex, East Sussex, UK

ABSTRACT Limited success in community-based conservation has challenged the role of institutions and design of incentive-based approaches. Earlier research has indicated the importance of a supportive institutional framework for the provision of effective incentives. This paper examines this issue for biodiversity conservation at Bardia National Park in Nepal. Derived from a wide survey of literature, seven criteria to assess institutional performance at Bardia were identified. Issues such as illegal resource extraction, park-people conflict, and deterioration of trust in NGOs indicated institutional failure. On the other hand, the park management has supported local communities by forming financial savings groups, and establishing a buffer zone management council to build social capital. This has resulted in increased collective action, greater trust within communities, enhanced citizen trust in the park management and more positive attitudes towards biodiversity conservation. Findings presented here indicate that even though the initial design of conservation initiatives often includes some of the characteristics necessary for success, these tend to be lost in the implementation process, with individuals in positions of responsibility shaping the process to suit their interests. Currently, the State acts as a mediator between the local communities and NGOs. This research suggests that instead, the communities themselves should be given a central position between the State and the NGOs, with all programmes channelled through them. By enhancing technical capabilities, programme monitoring, and building of institutions through trust, cooperation and collective action, such a rearrangement of roles would contribute not only to an effective design of conservation initiatives, but also to their effective implementation.

Complimentary Contributor Copy

106

Shova Thapa Karki

Keywords: Institutions; incentives; community-based conservation; community-based natural resource management; institutional performance

INTRODUCTION Historically, protected areas in developing countries have been set up according to Western approaches, often labelled as “fortress conservation” or “fences and fines”, whereby locals are excluded from the designated area, and the authority to manage natural resources is given to the central government (Barrett et al., 2001; Hulme and Murphree, 2001). Management relying on this mechanism has failed in enforcing effective rules for biodiversity conservation, contributed to vulnerability of locals by displacing them from their homeland, and threatened the biodiversity that it was supposed to protect (Barrett and Arcese, 1995; Brockington, 2004). Furthermore, the exacerbation of conflict between resource management institutions and local communities in these situations frequently diminishes local support, thereby compromising the chance of achieving the conservation objectives (Brown, 2003). Recognising the limits and failures of such management institutions, conservation initiatives1 have started to incorporate various incentive programmes to enhance collaborative management. An incentive for biodiversity conservation is here defined as “a specific inducement designed and implemented to influence government bodies, business, nongovernmental organisations or local people to conserve biological diversity or to use its components in a sustainable manner" (Emerton, 2000: 2). By diverting land, capital and labour towards conserving biological resources, incentives help to promote broader participation in conservation that will benefit not only resources but also rural communities (McNeely, 1988). Three kinds of incentives are commonly in use: 1) direct incentives in cash or in kind such as tax breaks, grants, compensation for animal damage, and differential access to resources; 2) indirect incentives setting enabling conditions such as building communitylevel conservation institutions, involving community in management and decision-making about resource use; and 3) disincentives in the form of penalties for non-compliance (McNeely, 1993; OECD). One of the most popular approaches to indirect incentives has taken the form of community-based conservation (CBC2), which allows local people to be involved in the management of natural resources, and aims to overcome interest-based conflicts by integrating conservation with economic development objectives (Selfa and Endter-Wada, 2008; Tai, 2007). CBC initiatives include monetary compensation, revenue-sharing schemes, and designation of buffer zone areas that allow local people access to resources, wildlife utilisation, and income-generating activities. These incentives are believed to help compensate for the substantial social and economic consequences that the establishment of protected areas may cause to these communities (West et al., 2006).

1

2

The term ‘initiative’, ‘approach’, and ‘measures’ in this paper refer to all benefit programmes used under conservation incentives. CBC has been broadly used to include other related concepts such as incentive-based conservation (IBC), community based natural resource management (CBNRM), and integrated conservation and development projects (ICDP). Although named different, these generally adopt a common development strategy that aims to reconcile conservation with the development needs of communities through benefit sharing.

Complimentary Contributor Copy

Community-Based Conservation

107

However, despite repeated efforts over the last two decades, these initiatives have been largely unsuccessful in integrating conservation and development (Barrett and Arcese, 1995; Brown, 2003); their performance has proven limited and unconvincing in many cases (Selfa and Endter-Wada, 2008; Spiteri and Nepal). Incentives have rarely targeted communities’ needs successfully, and benefits tend to accrue primarily to elites and higher social castes (Jones, 2007). Critics have accused CBC initiatives of being unable to provide equal benefits to the different rural communities affected (Spiteri and Nepal, 2008), for being socially unjust (Brockington et al., 2006), and for not protecting biological diversity (Chan et al., 2007). One of the key reasons for this lack of effectiveness lies in the institutional design of CBC programmes (Brechin et al., 2002; Brown, 2003; Tai, 2007). As Barrett et al., (2001) claim, CBC projects that are designed without due regard to broader institutional considerations at multiple levels will only serve to inflame park-people conflict. Similarly, biodiversity conservation incentives are effective only within a supportive institutional framework (Wells, 1998) involving institutions that are capable of implementing, monitoring, enforcing and assessing policies and regulations at local, national and international levels . These arguments highlight the crucial role of institutions, and the need for robust institutional design of CBC programmes and their sustainable management. Institutions are defined as “durable systems of established and embedded social rules that structure social interactions” (Hodgson, 2006:13), therefore goes beyond simple rules and incorporates social rule systems. Institutions include both formal and informal interactions between individuals and groups in society. Formal institutions generally apply formal rules, designed for a specific purpose, to specify those actions or outcomes that are permitted, prohibited or required, and to prescribe sanctions for rule violation (Crawford and Ostrom, 1999). These contrast with the informal institutional rules of a society, such as social norms, customary laws and codes of conduct, and their enforcement mechanisms (e.g. social networks) that together guide people’s behaviour (Ostrom, 1990). Informal institutions do not impose formal and legal sanctions on violators of such rules; instead, compliance is learned through repeated interactions within the society and emerge through day to day practice (Ostrom, 1990). The design, functioning and existence of institutions determine the manner in which resources are allocated and how the income derived through resource management is distributed (Dong et al., 2009). Ensuring such functionality requires evaluation of institutions (Bellamy et al., 2001). Evaluating institutions and institutional design can be a foundation for understanding incentives and their impacts on resource conservation (Brown, 2003). While it is important to evaluate the impact of incentive measures on community well-being, such evaluation cannot be carried out without due regard to the institutional performance (Wells, 1998). However, measuring institutional performance is a complex issue as it involves quantifying the performance of policies, regulations, norms and traditions (Balint et al., 2002). In practice, institutions do not ‘perform’, but rather affect the performance of conservation indirectly, for instance by influencing the effectiveness and efficiency of natural resource management (Balint et al., 2002). Balint et al., (2002) noted that only a few studies have made use of criteria to explicitly assess institutional performance in natural resource management. By developing a set of criteria for assessing institutional performance at a specific case study area in Nepal, this study therefore provides a first step towards the kind of indicator-based assessment called for by Balint et al.

Complimentary Contributor Copy

108

Shova Thapa Karki

As an early adopter of CBC, Nepal has implemented various conservation incentives in its protected areas, many of which fully recognise the local people’s participation in the management of park resources. However, low level of community participation in park management (McLean and Straede, 2003; Mehta and Kellert, 1998), negative attitudes towards the NGOs (Allendorf et al., 2007), dependency on the park resources for livelihood purposes and resource use conflicts (Baral and Heinen, 2007) have posed challenges for the role of institutions and compromised the effectiveness of CBC initiatives for biodiversity conservation and community well-being. Therefore, the aim of this paper is to assess the role of institutions and incentives in community well-being and biodiversity conservation at Bardia National Park (BNP), using some of the key characteristics of community-based conservation, drawn from existing literature on CBC, as an assessment criteria.

EFFECTIVENESS OF CBC INITIATIVES: ASSESSING INSTITUTIONAL PERFORMANCE Evaluating institutional performance requires identification and definition of a number of criteria against which the institution can be evaluated (Corbera et al., 2009). In this research, general criteria rather than specific criteria to the assessment of performance is applied. This is mainly because not all impacts are amenable to measurement by criteria and it is not possible to define a single criterion that covers the entire range of factors present; broader contextual factors (e.g. natural resource endowments, socio-economic and political settings) also exert a crucial influence on the success of institutions (Gottret and White, 2002). Keeping these limitations in mind, several studies were referred to identify characteristics of effective community-based conservation, and to specify those criteria that are specifically applicable to the assessment of institutions responsible for biodiversity conservation and management at BNP. Effective conservation initiatives, for this research is defined as those key characteristics that are critical in achieving biodiversity conservation while supporting communities’ social, economical and cultural needs sustainably in a variety of contexts.

Criteria Used for Evaluating Effective CBC Initiatives Different criteria have been used for evaluating institutional performance in natural resource management. For instance, Imperial (1999) proposes compliance, feasibility, policy outcomes, and transaction costs to assess the performance. Saleth and Dinar (1999) carried out a comprehensive evaluation of the institutional performance of water institutions with particular attention to the effectiveness of formal institutions (Saleth and Dinar, 1999). Ostrom identified 8 design principles to create robustness in institutions (1990), expanded later to a set of 33 design principles (Agrawal, 2002). However, these design principles do not consider the need for adaptability of institutions in dynamic conditions, which is a key requirement in CBC initiatives (Armitage, 2005). Kellert et al., (2000) used six social and environmental attributes of CBC initiatives to compare the performance of such initiatives on three different continents. Building on previously proposed attributes, (Sheppard et al., 2010) developed a framework covering three

Complimentary Contributor Copy

Community-Based Conservation

109

important principles of short and long term socio-economic and ecological variability, and linkage mechanisms that foster the continued integration of development and conservation, to assess community-governed environmental management in Ghana. Focussing on good governance outcomes of a protected area, (Lockwood, 2010) identified seven principles – legitimacy, transparency, accountability, inclusiveness, fairness, connectivity and resilience as a platform for the assessment of governance quality. Similarly, (Gruber, 2010) drew upon 23 scientific articles to identify 12 principles and key characteristics of effective conservation initiatives. These principles were “field tested” in 45 case studies submitted at a World Bank workshop on community based natural resource management. Gruber’s characteristics of effective conservation initiatives are more explicit than Kellert’s, Sheppard et al.’s, and Lockwood’s variables, and cover a wider range of characteristics. While all include equity, participation, empowerment, and conflict resolution, Kellert et al., have omitted other important characteristics such as monitoring, accountability, social capital, and ‘adaptive co-management’ whereas Gruber excludes biodiversity protection and sustainability. In addition, none of the list considers coordination, flexibility and learning as important characteristics. All these characteristics are critical for robust and resilient institutions (Baral et al., 2010) and institutional performance measurement must address how communities reorganise and evolve in the face of perturbations, thereby fostering learning (Armitage, 2005). Literature covering other areas in natural resource management such as water management (Balint et al., 2002), community forest (Pagdee et al., 2006) and rangeland management (Dong et al., 2009) were also examined. While Balint et al., focus on formal and specifically designed institutions for water management including water law, water policy, and water administration, Dong et al., also considers informal institutions along with formal institutions, such as community-based institutions and State institutional arrangements, respectively. Both research groups have stressed resource use conflict resolution, participation, accountability, transaction costs, and transparency as important criteria for measuring institution performance. Pagdee et al., (2006) identified nine important factors for the success of community forest management. On the basis of a review of 31 research articles and 69 case studies, Pagdee et al., stressed that participation, decentralisation, community features, incentives, financial, human and technical support, and physical features are important determinants of success. In summary, above cited literature review revealed that the performance of institutions was typically assessed in terms of their outcomes (e.g. distributional equity and empowerment), procedural mechanisms (e.g. conflict resolution and participation), enabling conditions (e.g. coordination and transparency) and ability to change (flexibility). This implies that these different criteria are linked together; for instance, procedural mechanism and enabling conditions facilitates outcomes, which allows for adaptability, therefore are all equally important for effective conservation initiatives. The literature review led to identify seven criteria that are likely to be relevant in evaluating how successful the institutions at BNP has been able to meet those criteria (Table 1). Although identified by other studies, these criteria are also relevant to Bardia because they include most of the core characteristics of effective CBC initiatives identified across a broad range of contexts, regions and resource management systems. They combine some of Kellert’s and Gruber’s characteristics for effective CBC initiatives, for instance monitoring and feedback; research and information development; communication and information dissemination; participatory decision-making; and participation and mobilisation.

Complimentary Contributor Copy

110

Shova Thapa Karki

Biodiversity protection and sustainable utilisation characteristics were not included as it is beyond the scope of this research, and the focus of this research was on community wellbeing. Outcomes such as empowerment and equity are crucial in supporting core value of conservation initiatives. Devolving power to local communities will empower community members to be responsible and take a greater role in environmental decision making (Armitage, 2005). Similarly, it has been argued that those conservation programmes that do not distribute the benefits in an equitable manner tend to lead to a “tragedy of the commons”, thereby making conservation objectives more difficult or impossible to achieve (Spiteri and Nepal). Table 1. Criteria, their definitions and assessment criteria used for assessing institutional performance

Ability to change

Enabling conditions

Procedural mechanism

Outcomes

Criteria

Definition used in the context of this research Empowerment The strengthening of local communities through an improvement in their role and responsibilities in relation to the park managers.

Equity

Conflict resolution

Assessment Criteria

Sharing of control and the distribution of decision-making rights to the communities, as well as capacity-building among women and the poor through their closer involvement in the preparation of park management plans and programmes. The equal distribution and allocation of Distribution of financial resources, salary, benefits and resources between different and the involvement of poorer households groups in a community. in income generating programmes. The conflict between local communities and park management resulting from problems caused by wildlife and the management rules. Meaningful inclusion and involvement of local people, e.g. is decision making, that goes beyond mere consultation.

The way compensation measures are handled and resolved with appropriate compensatory measures. Problems faced due the park. Participation Communities participating in planning, consulting, decision making and evaluating park management plans, and development projects. Transparency The degree to which information Local communities are clear about different and concerning the allocation of resources and project activities and their monitoring Accountability benefits is made publicly available. process. Different organisations working Accountability, refers to the degree to together to support community well-being which citizens can hold officials to and biodiversity conservation account for their actions and have access to information concerning decisionmaking. Coordination The synchronisation of the activities of Presence or absence of negotiation, different organisations towards a common monitoring and collaboration between goal, is important for negotiating, different organisations, working closely to monitoring and obtaining information support biodiversity conservation and about these activities. community development programmes. Flexibility and To adapt according to circumstances and Flexibility in changing or modifying rules adaptability to learn from previous experiences. according to shifting circumstances. E.g. alteration of rules, and learning from change

Complimentary Contributor Copy

Community-Based Conservation

111

Procedural mechanisms influences the outcomes of institutional performance so they need to be considered in evaluating institutions. For instance, conflict arises while trying to reconcile local livelihood needs with rules and regulation for biodiversity conservation. Central to mitigating such conflict is careful planning and development of strategies for conflict management (Spiteri and Nepal), which will also support empowerment and equity. Similarly, according to Campbell and Vainio-Mattila (2003) effective CBC must include appropriate public participation at all stages from information gathering to consultation, decision-making, initiating action and evaluation of the initiative programme. Engaging communities in a multi-stakeholder problem-solving and decision-making process can provide a forum for knowledge sharing and collaborative learning about biodiversity conservation, local needs, livelihoods, and the socio-economic system (Olsson et al., 2004). Effective communication, coordination and information exchange between all parties involved in biodiversity conservation, notably those who are affected by biodiversity conservation initiatives, fosters transparency, trust and information sharing that can be translated into usable knowledge for biodiversity conservation and community well-being (Gruber, 2010). Shared knowledge about different issues with regard to biodiversity conservation and conservation initiatives can support learning and adaptation in the community (Armitage, 2005). Furthermore, transparency is important for the sustainability of CBC, as a precondition for communities seeing CBC as a legitimate, trustworthy and credible organisation (Olsson et al., 2004). Linked social-ecological systems are constantly coevolving and adapting to changing local situations; therefore, management initiatives should also be flexible so as to adapt to new unforeseen situations (Berkes and Folke, 1998). The flexibility to adapt according to circumstances and to learn from previous experiences is crucial to the success of local management initiatives and their sustainability over the long term (Baral et al., 2010). Such flexibility can be gained by institutions through shifting responsibilities between hierarchies of management, altering rules so that they are more responsive to change, and increasing the human and social capital that facilitates learning from the change (Armitage, 2005; Baral et al., 2010).

MATERIALS AND METHODS Study area: Bardia National Park The Bardia National Park (BNP) is one of the largest in the lowland Terai, covering an area of 968 km2 in the Bardia and Banke districts in the western region of Nepal (see Figure 1). The climate is subtropical monsoon with a rainy season from June to early October, a cool dry season from October to late February, and a hot season from March to mid-June (Upreti, 1994). Seven major vegetation types have been found at BNP; four of them are forests: Sal forest, Khair-Sissoo forest, Riverine forest, Hard-wood forest; and three different grasslands: wooded grassland, Phanta and Tall floodplain grassland (Dinerstein, 1979). The park is home to endangered animals such as the Royal Bengal tiger (Panthera tigris), Asian elephant (Elephas maximus), Greater One-horned rhinoceros (Rhinoceros unicornis), swamp deer (Cervus duvauceli), and black buck (Antilope cervicapra).

Complimentary Contributor Copy

112

Shova Thapa Karki China

India

Bardia National Park

N

Protected Area Buffer Zone Study Area

Figure 1. Map of Bardia National Park and study sites: Thakurdwara, Shivapur and Suryapatuwa.

Altogether 20 Village Development Committees (VDCs), the smallest political unit in Nepal, border the national park. These villages fall within the buffer zone of the BNP. Three villages were selected as case study sites in the south-eastern part of the BNP, namely Thakurdwara, Shivapur and Suryapatuwa. The villages were selected on the basis of the following characteristics: i) location within the buffer zone boundaries, ii) large number of conservation initiatives in the form of development projects in the village, iii) provision of buffer zone community forestry and government forest, and iv) VDCs with highest wildlife densities (Wegge et al., 2009). Among the villages, Thakurdwara and Shivapur are also located close to the park boundary and Suryapatuwa further away. These villages consist of a mix of ethnic groups found in western Terai, dominated by Tharus, with Brahmins as the second largest group followed by Chettri, and other disadvantaged castes. Socio-economic charactersitics of the households at the three villages, are given in Table 2.

Complimentary Contributor Copy

113

Community-Based Conservation Table 2. Socio-economic characteristics of surveyed households at three villages Social attributes Total number of households Number of households interviewed Average age of the respondent Average household size Average land holding size per household (ha) Average livestock Average fuel wood consumption per household (kg/day) Average distance of households from the park (km) Residential Status of surveyed households (in percent) Households with local inhabitants Households with migrants Caste composition (in percent) Tharu Brahmin Chettri Disadvantaged Thakuri (higher caste) Other Wealth categories (in precent) Rich Medium Poor

Thakurdwara 1338 163 45.7±11.14 6.1±3.18 0.78±0.79 12.3±24.48

Shivapur 1219 144 48.0±12.50 7.8±5.2 0.65±0.59 13.5±11.86

Suryapatuwa 426 51 51.21±12.35 8.1±3.05 0.9±0.99 17.4±11.06

8.4±3.45 1.1±0. 9

10.31±7.41 0.6±1.1

10.26±5.34 2.4±0.9

44 56

9 91

45 55

46 26 12 10 2 4

36 25 23 13 2 1

40 14 16 12 6 12

9 49 42

6 29 65

24 43 33

Institutions and Incentives at BNP Different organisations have been involved in the management and conservation of biodiversity in Nepal. Under the Ministry of Forests and Soil Conservation (MFSC), the Department of National Park and Wildlife Conservation (DNPWC) coordinates and implements all the activities related to wildlife management in national parks, wildlife reserves and conservation areas. However, law enforcement within parks and reserves has been the responsibility of the Nepalese army since 1974. Policy formulation for wildlife conservation in protected areas is performed by the MFSC and the implementation of policies, rules, regulations and legislation is undertaken by the DNPWC. In 1994, CBC initiatives incorporating various incentives were introduced in the park in order to minimise park-people conflict and combine conservation and development goals (HMG, 2001). One of the immediate results was the establishment of a buffer zone and the corresponding development of regulations (in 1996) and guidelines (in 1998) (HMG, 1996). The buffer zone Act has made the provision to channel back 30-50% of the revenue generated from the park into community development activities (HMG, 1996). The objective of the buffer zone was to satisfy the subsistence needs of the local inhabitants residing in a defined

Complimentary Contributor Copy

114

Shova Thapa Karki

belt of park-impacted areas while increasing the area for wildlife habitat (HMG, 2001). Further incentives were introduced in the form of development projects such as direct and indirect employment as tourist guides or park staff, skills enhancement training, infrastructure support, development of health care and education, compensation payments, and implementation of development projects such as irrigation, vegetable farming, improved livestock breeding, agro-forestry, buffer zone community forestry, and annual grass-cutting programmes (HMG, 2001). These incentives are implemented through partnership between DNPWC and various national and international organisations (Brown, 2003). Among them the key stakeholders are different INGOs and NGOs.

Actors Involved, Their Role and Responsibilities and Interaction with Other Actors A range of actors are involved in the management of BNP that influence the communities’ well-being, figure 2. In the figure, Buffer Zone Management Council (BZMC) is the central point of interaction between the organisations and the communities.

NTNC, WWF, UNDP, CARE

Field office

Ministry of Forest and Soil Conservation

BZMC

DNPWC

BCP, TAL, PCP

BNP, Park office Social institutions (Badghar, kulopani, adhiya)

Communities

3

Figure 2. Actors involved in the management of the BNP and the communities .

All the development projects, discussions and negotiations channel via BZMC to the communities. Actors that have authority to set policies and to implement wildlife conservation measures at BNP are situated on the right side of figure 2. Other actors that have interests and funds to mobilise wildlife conservation are located on the top left side of figure 2. The main actors, their roles and responsibilities in park management and interactions with other actors are described in table 3.

3

NTNC: National Trust for Nature Conservation; WWF: Worldwide Fund for Nature; CARE: Cooperation for Assistance and Relief Everywhere; UNDP: United Nations Development Programme; BCP: Bardia Conservation programme of KMTNC; TAL: Terai-arc Landscape project of WWF; PCP: Participatory Conservation Programme of UNDP; BZMC: Buffer Zone Management Council; DNPWC: Department of National Park and Wildlife Conservation; BNP: Bardia National Park)

Complimentary Contributor Copy

Community-Based Conservation

115

Table 3. Actors involved, their role in park management and interactions with other actors Actors

Responsibilities

Bardia National Park The main actor with legal (BNP) authority for park (Ministry of Forest management. and Soil Conservation, DNPWC) NGOs and INGOs (FOs)4 (NTNC, WWF, PCP, CARE) Buffer Zone Management Council (BZMC)

Informal institutions (Badghar, Kulopani, Adhiya5, CBOs )

Role

Interaction with other actors Mediator between INGOs, Between FOs and NGOs and BZMC for BZMC implementing development projects. Maintain coordination between FOs. Facilitator in implementing Between BZMC development projects to the and DNPWC communities

Providing development projects in the form of economic and societal development. Disbursement of 30-50% of Locally formed formal Between revenue generated by park institutions. All the projects community-based for community channel via the BZMC organisations and development and above actors compensation. Collective action in the No role in park with the BZMC village, no responsibility management or for park management development projects

Data Collection Household Questionnaire Survey Altogether, 358 households (comprising 12 % of the total) were included in the study, using a random sampling strategy to select households from electoral registers. The questionnaires were usually addressed to household heads, who were mostly male. When these individuals were not available, the next willing member was interviewed. The questionnaire enquired about the households’ relationships with the park, through questions concerning four areas: i) the socio-economic background; ii) benefits received by the households from the park and its impact on their livelihoods, iii) interaction (meeting, support, help, participation, technical support) with officials from the park and other organisations, and iv) involvement in development projects and satisfaction with the programmes. All the questions were closed-ended, but households were given the opportunity to further elaborate on any issues they raised. Community Workshops and Focus Group Discussions Community workshops and focus group discussions were organised in each village to obtain information on the views of local residents on the impact of BNP on their livelihoods and well-being. These qualitative methods were designed to obtain an overview of the local people’s perspectives on conservation incentives, the role of the BNP and the importance of CBCs to their livelihoods. 4 5

Commonly called as Funding organisations (FOs) Badghar: village chief; Kulopani: water guard; Adhiya: sharecroppring; CBOs: Community-based organisations

Complimentary Contributor Copy

116

Shova Thapa Karki

For organising community workshops, VDC chairpersons were consulted, prior to the workshop, in order to explain the aim of the workshop. The village chairperson then notified households in each village about the workshop and encouraged them to participate. Altogether 30-35 people attended in each workshop, comprising different ethnicities, social classes and status, wealth groups, age groups and gender, and including teachers, political leaders and a chairperson of the buffer zone user’s committee. Five groups were formed with six or seven people in each group; each group was then given a list of issues to discuss: livelihood capitals, the level of dependency on park resources, factors that support or hinder livelihood, institutions, the role of development projects, and the role of the buffer zone community forest (BZCF) in fulfilling citizens’ needs. Groups were given an hour to discuss their specific topic, while the facilitators went around to observe the discussions. At the end of the discussion time, each group nominated a leader to present the group’s discussion to the other groups. This part was especially helpful, since other groups could fill in any gaps and omissions by the individual presenters. In each village, the entire session lasted four hours and participants contributed with full engagement. Twelve focus group discussions in the three villages were organised, with usually five or six members in a group, to discuss the above issues. These focus groups were organised whenever people were gathered for their own monthly savings meetings, forest user group meetings and when people were visiting the projects. This was time-efficient for the participants and encouraged them to speak freely amongst their usual group members.

Semi-Structured In-Depth Interviews To explore park management’s perspective, semi-structured in-depth interviews were carried out with officials from the park, and two NGOs that had been working in Bardia for some decades. In total, 11 officials from different organisations such as the park office, international and national non-governmental organisations and the buffer zone management council were interviewed. Three sets of interview questionnaires were developed, one set for the park office (including the park warden, park ranger and conservation officer at the central office), another for NGO programme officers, and a third for the council’s chairperson and user committee chairs. Subjects addressed by the questionnaire included park- people conflicts, conservation incentives and the aims of the conservation incentives, communities’ involvement in park management and the level of interaction with different organisation, development project distribution, mechanism of monitoring and working with other organisations.

Data Analysis Interviewees came from diverse backgrounds with different roles in the park management. Due to their different roles, such as project manager and the Buffer Zone Management Council chairperson, they had different understandings and explanations of people's behaviour and had different interests and needs from the conservation programme. Furthermore, they interacted differently and produced different types of ideas and meanings for conservation and the role of the communities in park management. Although use of keywords addressed cognition, affect, and interaction, it failed to capture the emerging stories

Complimentary Contributor Copy

Community-Based Conservation

117

of the meaning of the experience for each interviewee. Considering these issues, narrative analysis combined with simple descriptive analysis was taken for data analysis purposes. Narratives are a story relating to a series of events. “Narrative rationality understands synoptically the meaning of a whole, describing and integrating the parts making one meaning; and stresses that narrative accounts have a unique explanatory power" (Polkinghorne 1988 in Richards, 1989:258). Narrative accounts are courses of action rather than static variables; understanding the phenomena under consideration, and gathering scattered information and explaining why certain actions have taken place or how certain interactions happened. Therefore, narrative analysis is useful to understand how people perceive and conceptualise actions.

RESULTS AND DISCUSSION Empowerment: Devolution of Power The respondents, including the chairpersons of the buffer zone management council (BZMC) and the users committee (UC), argued that the buffer zone management guidelines had created tensions between the user groups of the buffer zone community forest (BZCF) and the BZMC. For example, the BZMC has greater control than the BZCF over the mobilisation of financial resources, i.e. park revenues for community development activities. Dissatisfaction with the rules was expressed by all the user group committee chairpersons in interviews, in the community workshop and in the focus group discussions. Complaints primarily concerned the asymmetries of power in the management guidelines, which had created tension between the chairs of the UC at BZCF and the BZMC. As both of these chairs had equal status but existed in different places, they wanted to control their natural and financial resources independently. However, having equal power at two different institutions, each wanted funds to pass through them; this had not only increased the administrative cost, but also delayed the flow of funds into the community-based organisations and user groups. This had already created collective action problems such as delays in the harvesting of community forest, little household involvement in road development programmes, and nonparticipation of forest user groups in community canal improvement. In the long run, this could hamper the implementation of development projects. Problems with regard to devolution of power have also been evident in other protected areas of Nepal; Heinen and Mehta (1999) have also observed that in conservation areas, power struggles occur between local institutions due to overlapping jurisdictions and mandates. The BZMC was given legal status so that control could be shared with the local communities, which could take responsibility for some of the decision-making on revenuesharing and the distribution of benefits. However, to date the park warden has had full control of these responsibilities within the BZMC, and consequently the local communities have claimed that they have no influence. This has created tensions on many occasions, for example the UC chairperson has accused the BZMC chairperson of concentrating all funds to his own village. This again resonates with the findings by Heinen and Mehta (2000), who found that the managerial structure of the buffer zone had remained largely top-down, with the park warden retaining additional powers in comparison to the buffer zone chair, stating

Complimentary Contributor Copy

118

Shova Thapa Karki

that park managers can dissolve the group and also denounce the community forestry status if the forest user groups do not act according to their rules. When asked about involving communities in decision-making, park managers conceded that this had not happened in practice; communities were consulted frequently to discuss problems and specific conservation issues, but were not given any decision-making power. Similarly Dong et al. (2009) found that decision-making, policy implementation and programme development at the national and district level are largely top-down processes in Nepal, often ignoring the needs and aspirations of local people. As articulated by Selfa and Endter-Wada (2008), “what is actually devolved to the local level is the implementation of decisions made elsewhere” (Selfa and Endter-Wada, 2008: 862). Empowerment also means improving the inclusion and representation of marginalised groups (Gruber, 2010). However, women and low-caste people were highly underrepresented within the executive committees of the BZMC. Furthermore, the only female participant in the BZMC’s annual meeting did not interact with the groups but merely listened to what others had to say during the entire two-day meeting. It is therefore uncertain whether the discussions had any relevance to her, or whether she understood them. Kellert et al., (2000) made a similar observation in two conservation areas of Nepal and Kenya, in which development projects purporting to build institutional capacity rarely included women and the more disadvantaged groups of society. Buffer zone management regulations have more recently made it mandatory to include women and disadvantaged groups in User Groups (Heinen and Shrestha, 2006), yet the representation of women in the UC and their involvement in buffer zone meetings remained limited in the present study.

Equity: Distribution and Allocation of Benefit Sharing Fifty-two percent of the households surveyed in the current study claimed that the project benefits had been distributed unequally, with higher castes and the otherwise wealthy and influential groups often benefiting more than the poor and vulnerable groups. For instance in Bardia, the high capital needs of biogas plants had prevented poor and vulnerable groups from participating in the biogas project. Similarly, there were other income generating projects where a participating household had to contribute a part of capital, which usually exceeded what a poor household could afford. Such limitations on the poor households’ ability to participate in a majority of the development projects at Bardia could engender negative attitudes towards the park and park programmes and between the communities as also stressed by Kellert et al. (2000) and Jones (2007). This may not only diminish support of the park management, but also create tension between communities that degrades social capital within these households. Wage distribution is also often unequal within a single organisation, which can generate dissatisfaction among employees (Gibson et al., 2005). This was evident in Bardia, where people undertaking similar jobs were paid different salaries by the government and an international NGO (INGO), causing potential conflict between the employees. Government officials also tend to favour higher-paid jobs over those that offer more responsibilities but a lower salary. An example of this is the involvement of BZMC assistants undertaking paid NGO work alongside their main job at the BZMC. The responsibility of these assistants is to provide necessary support for communities; however, they are often absent from this work in

Complimentary Contributor Copy

Community-Based Conservation

119

order to conduct their NGO jobs. This has created negative attitudes towards the NGOs and their programmes. Attitudes towards the protected areas are in general closely linked with protected area activities, park management, and the nature of interactions between locals and the staff of organizations responsible for providing incentive programs (Akama et al., 1995). Therefore, negative attitudes towards these organisations can hamper a community’s long term participation in conservation programmes through a reduction in collective action, erosion of trust in authorities, and an overall loss of community support.

Conflict Resolution: Addressing Park-People Conflicts Different direct compensatory measures are employed alongside with incentive measures in protected areas such as Bardia. These include cash compensation for human injury and property and livestock damage, timber collection to renovate houses damaged by wildlife, and an annual grass-cutting programme (GCP). The GCP is one of the highly commended programmes, which has to some extent fulfilled the needs of the communities. Respondents criticised the lack of monetary compensations for their lost crops and livestock due to park wildlife; and were unsatisfied with the way their complaints had been handled (See Table 46). Table 4. Number of compensation claims, actions taken and the opinions of households on the adequacy of the measures at Thakurdwara and Shivapur Compensation handling process Thakurdwara Shivapur Complaints made to the Park Office (%) 11 8 Adequacy of the measures taken in response to the3 9 complaints* (%) Actions taken by the Park Office on the complaints* (%) 18 0 Adequacy of the action taken to compensate damage (%) Inadequate 48 39 Adequate 7 0 Reasons for inadequacy of the measures (%) Lengthy process 58 61 Only few get compensated 35 43 Note: * = Percentage given corresponds to the households agreeing that adequate measures were taken. Similarly, households agreeing that actions were taken by the Park office on the complaints made by the households.

Park managers agreed that making complaints and taking actions was a lengthy process as they had to follow certain rather time-consuming procedures, such as sending the staff to make an observation, quantifying and valuing the damage, documenting the findings, and preparing a report, before the complaint could be processed. However, park management claimed that for human injury and livestock damage (mostly to cattle) they had handled compensation as a priority and provided for most of the damage. With regard to crop damage compensation, they highlighted difficulties in quantification, stressing that households often exaggerated the damage; hence crop damage compensation had not been possible. 6

Only households close to the park, i.e. Thakurdwara and Shivapur, but not Suryapatuwa, were affected by wildlife problems, and made complaints.

Complimentary Contributor Copy

120

Shova Thapa Karki

Participation: Role of Communities in Park Management The fourth amendment of the Nepalese Wildlife Conservation Act of 1973, introduced in 1994, stipulates participation of local people as one of the main aspects of park management in all of the protected areas. However, in Bardia, the participation of local people represented merely one-way communication, as the locals were involved in information collection only. They were not allowed to participate in decision-making at a higher strategic policy level, and had rarely been given access to decision-making concerning the park management rules. For example, the chairman of BZMC claimed that he had never been invited to the monthly meetings between the project coordinators of different INGOs and NGOs and the park warden. Despite the rhetoric of the importance of participation in the discourse of park managers and project coordinators of funding organisations, the issue was rarely mentioned by the households interviewed and observed in this study. Similar observations have been made in other protected areas, indicating that while communities may be involved in the planning of incentive programmes, they are often given only limited opportunities for partnerships and active participation (Heinen and Mehta, 1999; Kellert et al., 2000). Households were also reluctant to participate in meetings, which they did not believe would benefit them. Even during the focus group meetings organised for data collection for this study, only higher caste groups participated actively. According to the project coordinators at Bardia, most visitors to the park office were among those who had been provided with incentives and had benefited from such incentive programmes. Visits by women were very rare, presumably because women felt too nervous to talk to the predominantly or exclusively male park office staff. One of the main objectives of participation by local people is to incorporate their traditional knowledge into the management of the natural resources on which the locals depend upon for their survival. However, such objectives had rarely been pursued in the BNP. Informal institutions can form the backbone of a broader understanding of issues in development and conservation (Spencer and Guzinsky, 2010). Therefore, park management plan needs to be designed in such a manner that it takes into account the influence of informal institutions, namely the constraints and opportunities that they pose for the implementation of the plan. Only recently, some steps have been taken to involve the village chiefs of buffer zone villages in park programmes at Bardia. However, the incorporation of the village chiefs and their traditional knowledge is still rare in most of the park management activities, such as the elaboration of a park management plan, the annual meeting of the buffer zone management council, and the project coordinator meetings. An example of the importance of local, traditional knowledge came from observations made by Tharus, one of the indigenous groups of the Nepalese Terai region with a long historical experience and knowledge on the plants and forests of the region. In Bardia, many of the local healers had been using different plant species for medicinal purposes for generations, while some of the elderly also had knowledge of specific fruits and vegetables that are found only in some parts of the park. There were still people who remembered their migration from the hills to the Bardia forest bringing their cattle herds to the Phanta (open grassland, which is the part of the park now and hold highest population). These people mentioned that they only travelled during specific months of the year to allow the grass to regenerate. During their temporary stay in the grasslands, they had also met other Indian herders bringing their cattle. As food was limited and they had to stay for several months,

Complimentary Contributor Copy

Community-Based Conservation

121

they used to go inside the park and find seasonal wild fruits, vegetables and plants to supplement their diet. According to one of the respondents, “I was only a child when I accompanied my father. There were plenty of grasses for everybody. We could eat lots of fruits and vegetables that I had never seen in my village. If you fell and had a wound there were also plants to use as antiseptics. Once we were here, we never went anywhere for three months and we did not have any problems. There were lots of deer species and also large numbers of them; we did not kill them as it was important to have them around for the grasses to grow.” (Thankurdwara village)

Such knowledge is highly important in addressing conservation of wildlife and fulfilling a community’s needs; however not much effort has been made to use that valuable knowledge. Sharing of local knowledge and integrating it into the park management may help in fostering social learning. Social learning is a process that occurs when people engage with each other, sharing their ideas and experiences to develop a common understanding as a basis of joint action. As argued by Webler et al., (1995), social learning is important in natural resource management and environmental decision-making as it can strengthen public participation and support collective action, reflecting collective needs and understandings. Further, social learning can contribute to decision-making by creating new cooperative relationships, which support resource management and transform adversarial relationships (Schusler et al., 2003).

Flexibility and Adaptability: Institutional Resilience During the formation of BZCF at Bardia, different user groups, including all of the households, user committees and chairpersons, were also formed for the management of the forest, and to establish rules for the harvesting and collection of non-timber forest resources, grazing inside the forest, and timber collection. However, the user groups of BZCF did not have the right to make decisions or to modify the existing rules of the BZCF. Respondents commented that even for daily harvesting and collection of resources they had to obtain permission from the park warden. The BZCF guidelines mention devolution of power to the communities, yet in reality such devolution exists only on paper and the park warden still has sole control over the rules. As the user groups have devised a mechanism for sustainable allocation and utilisation of resources, the park management could provide some decision making rights to households such as deciding who to exclude from the group, when to collect the resources, and at which plot of the forest grazing should be allowed creating flexible institutions. This would still allow the park warden to retain control over major decisions, such as timber harvesting. Despite restoring the BZCF through plantation and conservation, and managing it, forest user groups do not have any power to change the rules that were devised during the formation of the BZCF. Although the BZCF has replicated the community forest (CF) model, it still follows traditional protected area management practices by essentially restricting devolution of power to the communities. As noted by Nagendra et al., (2004) despite supporting forest generation, limited downward accountability, constraints on function, and a low level of effective control over forest-management policies could have

Complimentary Contributor Copy

122

Shova Thapa Karki

significant negative impacts on future programmes, causing an abstinence of local participation in the BZCF. Example form Shivapur village observed during the survey provides evidence to what Nagendra et al., forecasted. Unlike the two other villages, Shivapur does not have a BZCF or a CF for resource collection. However, it has the highest percentage (65 %) of households involved in illegal resource extraction from the park. Consequently, Shivapur villagers are frequently in conflict with the park management, and to resolve the problem, the park management offered to convert Shivapur Phanta (which is currently in a degraded state) to a BZCF. However, local households rejected the proposal, as explained by participants in our community workshop: “The park management wants free labour; that is why they are giving us that degraded forest. Once we restore it and there is sufficient regeneration, they will take over the control and authority. We will have no control of anything. So why would we regenerate that degraded forest and then later let them enjoy.” (Participants at community workshop in Shivapur village)

Shivapur Phanta was provided to only two wards of Shivapur village, with other wards not having access to community forest. When park managers were asked about providing some parts of the park as a BZCF, they rejected the idea, arguing that if they applied this rule to Shivapur village they would have to make similar offers to the other buffer zone villages. By stressing that the same rules must apply to all cases, they neglected the site-specific context. This is a typical example of a ‘panacea’, referred to by Ostrom et al. (2007), where a management system developed for one case and one system is applied everywhere across all user groups, reducing the flexibility to adapt to changing circumstances, and thus increasing the vulnerability of the ecological and social system. While it is true that it is necessary by law to apply the same rules across villages in order to ensure equal treatment of citizens, these concrete management solutions should be flexible enough to allow site-specific context to be taken into account. In the case of Bardia, for instance, the park management should consider restoring or using some areas of degraded forest such as Phanta in areas where these exist, and using road-side plantations to provide some funds to mobilise the villages in others.

Coordination: Knowledge and Information Sharing One of the roles of the park officer is maintaining coordination between the different NGO programmes. According to the project coordinators, this is ensured through a monthly meeting between all the project coordinators and the park warden, aimed at facilitating joint project implementation and avoiding the duplication of work. However, during the community workshop, communities and even the chairman of the BZMC reported a lack of coordination and a duplication of work between the organisations. Communities from the Shivapur village argued that rather than coordinating and supporting the projects, NGOs were competing between each other in an attempt to portray themselves as highly effective and efficient. Further, NGO officers stressed that in principle they were due to meet once a month to share information, but that in reality this had not taken place for the past six months. When asked about programmes on which the NGOs had collaborated, the officers argued that since

Complimentary Contributor Copy

Community-Based Conservation

123

each programme was unique, they had never interfered with other programme activities. When the park officers were asked about their role in the coordination process, they also mentioned the monthly meeting but did not give examples of any activities where they had worked together with other NGOs. NGO officers also mentioned that to avoid overlapping of programmes, they do not work in the same village. As all the NGO programmes have to be channelled through the park office, our observations during the community workshops suggested that the park office did not adequately monitor the NGO activities. For example, households from Shivapur village complained that one of these NGOs had chosen only two households for the installation of solar panels. During our community workshop, participants were still angry about how the money invested for those two houses could have been better invested for the benefit of the entire village by buying lamp posts, electricity wires and transformers. This perceived bias had generated negative attitudes among the households towards the NGOs. This may not only hinder community participation in development projects but can also degrade trust among the NGO officials.

Transparency and Accountability: Allocating Resources According to the communities observed and interviewed at Bardia, the project implementing and funding organisations lacked transparency and accountability. The openness of these organisations was perceived by the villagers as insufficient; the villagers did not know what areas the organisations worked in, how much funding they were planning to mobilise, and on which projects. The communities also claimed that the NGO staff never came to see whether there was an actual need for a project, or whether the projects mentioned in the operational plan for the communities were appropriate and potentially sustainable. They also mentioned that the NGOs worked according to their own criteria regardless of whether there was a need for the project or not. According to one respondent during the community workshop: “Different projects are working in their own areas without considering the need of communities. They measure the impact based upon how many biogas installations they have supported or how many households they have provided with alternative incomes. They use these numbers to demonstrate their success over the other organisation and to justify the funding they receive. There is no monitoring of the activities or evaluation of whether the benefit has gone to the household that needs it most.” (Shivapur VDC)

Because neither households nor the BZMC knew where the NGOs focussed their fund mobilisation, developed proposals rarely fit in terms of development priorities within the community interests. Due to the lack of clarity in the development, implementation and distribution of incentive programmes, and the small number of households chosen to receive the benefits, communities were not only losing trust in the NGOs, but they also often did not realise that they had actually received benefits. This perception regarding incentive measures is partly rooted in historical mistrust between local communities and conservation agencies (Spiteri and Nepal). Mehta and Kellert (1998) also highlighted problems of transparency and

Complimentary Contributor Copy

124

Shova Thapa Karki

accountability in Nepal’s Makalu-Barun Conservation Area and observed that sectoral offices did not implement all of the programmes mentioned and prioritised in their plans.

CONCLUSION AND IMPLICATIONS Derived from prior literature on different natural resources and community based conservation initiatives, seven criteria provided a framework for assessing CBC programmes, which was then applied to the case study of Bardia. Two major conclusions can be drawn on the usage of institutional design principles – one of a theoretical nature and the other more practical. At the theoretical level, it was found that the seven criteria share a common vision and objective. Empowerment, equity, conflict resolution, participation, flexibility and adaptability, coordination, transparency and accountability are all aimed at bringing local communities into the centre of park management and NGOs. This would devolve power, give local communities a stake in decision-making, and help share information about park management, therefore making communities more responsible. This should not only foster learning between all the stakeholders, but would also be a key to building social capital as an outcome of the learning process triggered by those criteria. It would also be likely to enhance technical capabilities, monitoring of the programmes, and institution-building through trust, cooperation and collective action, not only for effective design but also for effective implementation of conservation initiatives. Equity sharing, participation, empowerment and prioritising communities' needs have been well documented in the BNP management guidelines, and also adopted up to a point in practice. Institutions and institutional arrangements, including the park office and NGO programmes, sometimes empowered local communities, and supported their development activities. Given that households show concerns, particularly on the NGOs, with regards to equality of benefit distribution, and coordination and transparency, these need further attention. Favouritism and unequal benefit sharing have already resulted in negative attitudes towards the project implementing organisations. Institutions at Bardia were particularly poor at coordination between organisations as well as in ensuring the transparency and accountability of the programmes, although some efforts had been made to improve performance on these criteria. Wells (1998) argued that to improve an understanding of incentives for biodiversity conservation there is a need of case studies providing specific examples and information on different institutional framework, socio-economic conditions and ecological context. He further argues that institutions are important for biodiversity conservation as institutional failure contributes to biodiversity loss. Institutions fail when they are poorly designed, do not coordinate effectively with other institutions and do not exist (Wells, 1998). Going further, Barret et al., (2001) stress that successful conservation institutions must possess authority, ability to restrict access and use, technical capacity to monitor and managerial flexibility to alter incentives and rules of access. Many studies have highlighted the lack of empowerment, involvement, equity, conflict and accountability of institutions for biodiversity conservation in various case study settings improving our understanding of current state of art and justifying why these need to be considered for biodiversity conservation in protected areas.

Complimentary Contributor Copy

Community-Based Conservation

125

This research went one step forward adding the importance of institutional flexibility and adapatability, and of coordination between different actors to support biodiversity conservation institutions. To enhance empowerment, equity and participation, coordination between different organisations and monitoring from the central/governmental organisations is important. While assessing institutional performance this research did not analyse the role of community level institutions in biodiversity conservation. Investigating these roles could help in understanding what community can do in their part and how different organisations can strengthen the community for biodiversity conservation and their well-being.

ACKNOWLEDGMENTS I acknowledge Prof. Sigrid Stagl and Dr. Markku Lehtonen contribution in providing feedback at various stages during the preparation of this chapter. Fieldwork was supported by the Rufford Small Grant programme, UK, the Royal Geographical Society, and the World Wildlife Fund, Nepal. Thanks are also to the assistants for conducting the survey and to all the villagers for agreeing to participate. Finally, thanks goes to Prof. Jouni Paavola for reviewing the article and providing valuable feedback.

REFERENCES Agrawal, A. (2002). Common Resources and Institutional Sustainability. In The Drama of Commons: Committee on the Human Dimension of Global Change. E. Ostrom, Dietz, T., Dolsak, N., Stern, P.C. Stovich, S., Weber, E.U. (Eds.), National Academy Press, Washington D.C.: 41-86. Akama, J. S., Lant, C. L. and Burnett, G. W. (1995). Conflicting Attitudes toward State Wildlife Conservation Programs in Kenya. Society & Natural Resources 8(2): 133-144. Allendorf, T. D., Smith, J. L. D. and Anderson, D. H. (2007). Residents' Perceptions of Royal Bardia National Park, Nepal. Landscape and Urban Planning 82(1-2): 33-40. Armitage, D. (2005). Adaptive Capacity and Community-Based Natural Resource Management. Environmental Management 35(6): 703-715. Balint, B. E., Forkutsa, I. and de Freitas, A. C. R. (2002). Indicators for the Measurement of Institutional Performance Concerning Water Management. Application for Uzbekistan and Ghana. Center for Development Research, University of Bonn, http://www.pasad.uni-bonn.de/indicators_of_institutional_performance.pdf Accessed on 20.08.2010. Baral, N. and Heinen, J. T. (2007). Resources Use, Conservation Attitudes, Management Intervention and Park-People Relations in the Western Terai Landscape of Nepal. Environmental Conservation 34(1): 64-72. Baral, N., Stern, M. J. and Heinen, J. T. (2010). Growth, Collapse, and Reorganization of the Annapurna Conservation Area, Nepal: An Analysis of Institutional Resilience Ecology and Society 15(3): [online] URL: http://www.ecologyandsociety.org/vol15/iss13/art10/.

Complimentary Contributor Copy

126

Shova Thapa Karki

Barrett, C. B. and Arcese, P. (1995). Are Integrated Conservation-Development Projects (Icdps) Sustainable? On the Conservation of Large Mammals in Sub-Saharan Africa. World Development 23(7): 1073-1084. Barrett, C. B., Brandon, K., Gibson, C. and Gjertsen, H. (2001). Conserving Tropical Biodiversity Amid Weak Institutions. Bioscience 51(6): 497-502. Bellamy, J. A., Walker, D. H., McDonald, G. T. and Syme, G. J. (2001). A Systems Approach to the Evaluation of Natural Resource Management Initiatives. Journal of Environmental Management 63(4): 407-423. Berkes, F. and Folke, C. E., Eds. (1998). Linking Social and Ecological Systems : Management Practices and Social Mechanisms for Building Resilience., Cambridge: Cambridge University Press, Cambridge. Brechin, S. R., Wilshusen, P. R., Fortwangler, C. L. and West, P. C. (2002). Beyond the Square Wheel: Toward a More Comprehensive Understanding of Biodiversity Conservation as Social and Political Process. Society & Natural Resources 15(1): 41-64. Brockington, D. (2004). Community Conservation, Inequality and Injustice: Myths of Power in Protected Area Management. Conservation Society 2: 411-432. Brockington, D., Igoe, J. and Schmidt-Soltau, K. (2006). Conservation, Human Rights and Poverty Reduction. Conservation Biology 20(1): 250-252. Brown, K. (2003). Integrating Conservation and Development: A Case of Institutional Misfit. Frontiers in Ecology and the Environment 1(9): 479-487. Campbell, L. M. and Vainio-Mattila, A. (2003). Participatory Development and CommunityBased Conservation: Opportunities Missed for Lessons Learned? HUMAN ECOLOGY 31(3): 417-437. Chan, K. M. A., Pringle, R. M., Ranganathan, J., Boggs, C. L., Chan, Y. L., Ehrlich, P. R., Haff, P. K., Heller, N. E., Al-Krafaji, K. and Macmynowski, D. P. (2007). When Agendas Collide: Human Welfare and Biological Conservation Conservation Biology 21: 59-68. Corbera, W., Gonazalez Soberanis, C. and Brown, K. (2009). Institutional Dimensions of Payment for Ecosystem Services: An Analysis of Mexico's Carbon Forestry Programme. Ecological Economics 68: 743 - 761. Crawford, S. and Ostrom, E. (1999). A Grammer of Institutions. American Political Science Review 89: 582-600. Dinerstein, E. (1979). An Ecological Survey of the Royal Karnali-Bardia Wildlife Reserve, Nepal. Part I: Vegetation, Modifying Factors, and Successional Relationships. Biological Conservation 15(2): 127-150. Dong, S. K., Lassoie, J., Shrestha, K. K., Yan, Z. L., Sharma, E. and Pariya, D. (2009). Institutional Development for Sustainable Rangeland Resource and Ecosystem Management in Mountainous Areas of Northern Nepal. Journal of Environmental Management 90(2): 994-1003. Emerton, L. (2000). "Using Economic Incentives for Biodiversity Conservation." Retrieved http://data.iucn.org/dbtw-wpd/edocs/PDF-2000-002.pdf 19.07.2003. Gibson, C. C., Andersson, K., Ostrom, E. and Shivakumar, S. (2005). The Samaritan's Dilemma: The Political Economy of Development Aid, Oxford University Press. Gottret, M. V. and White, D. (2002). Assessing the Impact of Integrated Natural Resource Management: Challenges and Experiences. Conservation Ecology 5(2): art. no.-17.

Complimentary Contributor Copy

Community-Based Conservation

127

Gruber, J. S. (2010). Key Principles of Community-Based Natural Resoruce Management; a Synthesis and Interpretation of Identified Effective Approaches for Managing the Commons. Environmental Management 45: 52-66. Heinen, J. T. and Mehta, J. N. (1999). Conceptual and Legal Issues in the Designation and Management of Conservation Areas in Nepal. Environmental Conservation 26(1): 21-29. Heinen, J. T. and Mehta, J. N. (2000). Emerging Issues in Legal and Procedural Aspects of Buffer Zone Management with Case Studies from Nepal. Journal of Environmental & Development 9( 1): 45-67. Heinen, J. T. and Shrestha, S. K. (2006). Evolving Policies for Conservation: An Historical Profile of the Protected Area System of Nepal. Journal of Environmental Planning and Management 49(1): 41-58. HMG, N. (1996). Buffer Zone Management Regulation (Unofficial Translation). K. Department of National Parks and Wildlife Conservation, Nepal HMG, N. (2001). Royal Bardia National Park: Management Plan 2001-2005, His Majesty's Government of Nepal. MInistry of Forests and Soil Conservation, Department of National Parks and Wildlife Conservation Hodgson, G. M. (2006). What Are Institutions? From Orders to Organisations. Journal of Economic Issues 40(1): 1-25. Hulme, D. and Murphree, M. (2001). African Wildlife and Livelihoods. The Promise and Performance of Community Conservation, James Currey Ltd. Imperial, M. T. (1999). Institutional Analysis and Ecosystem-Based Management: The Institutional Analysis and Development Framework. Environmental Management 24(4): 449-465. Jones, S. (2007). Tigers, Trees and Tharu: An Analysis of Community Forestry in the Buffer Zone of the Royal Chitwan National Park, Nepal. Geoforum 38(3): 558-575. Kellert, S. R., Mehta, J. N., Ebbin, S. A. and Lichtenfeld, L. L. (2000). Community Natural Resource Management: Promise, Rhetoric and Reality. Society and Natural Resources 13(8): 705-715. Lockwood, M. (2010). Good Governance for Terrestrial Protected Areas: A Framework, Principles and Performance Outcomes. Journal of Environmental Management 91: 754766. McLean and Straede, S. (2003). Conservation, Relocation, and the Paradigms of Park and People Management - a Case Study of Padampur Villages and the Royal Chitwan National Park, Nepal. Society and Natural Resources 16(6): 509-526. McNeely, J. A. (1988). Economics and Biological Diversity: Developing and Using Economic Incentives to Conserve Biological Resources. Gland, Switzerland, IUCN. McNeely, J. A. (1993). Economic Incentives for Conserving Biodiversity-Lessons for Africa. Ambio 22(2-3): 144-150. Mehta, J. N. and Kellert, S. R. (1998). Local Attitudes toward Community-Based Conservation Policy and Programmes in Nepal: A Case Study in the Makalu-Barun Conservation Area. Environmental Conservation 25(4): 320-333. Nagendra, H., Karmacharya, M. and Karna, B. (2004). Evaluating Forest Management in Nepal: Views across Space and Time. Ecology and Society 10(1): 24. OECD (1994). Economic Incentive Measures for Conservation and Sustainable Use of Biological Diversity: Conceptual Framework and Guidelines for Case Studies. Environmental Monograph No.97. Paris, OECD.

Complimentary Contributor Copy

128

Shova Thapa Karki

Olsson, P., Folke, C. and Berkes, F. (2004). Adaptive Comanagement for Building Resilience in Social–Ecological Systems. Environmental Management 34( 1 ): 75 - 90 Ostrom, E. (1990). Governing the Commons: The Evolution of Institutions for Collective Action, Cambridge: Cambridge University Press. Ostrom, E., Janssen, M. A. and Anderies, J. M. (2007). Going Beyond Panaceas. Proceedings of the National Academy of Sciences of the United States of America 104(39): 1517615178. Pagdee, A., Kim, Y. and Daugherty, P. J. (2006). What Makes Community Forest Management Successful: A Meta-Study from Community Forests through the World. Society & Natural Resources 19: 33-52. Richards, R. J. (1989). Narrative Knowing and the Human Science. The American Journal of Sociology 95(1): 258-260, Book Review. Saleth, R. M. and Dinar, A. (1999). Evaluating Water Institutions and Water Sector Performance, World Bank Technical Paper No. 447. Schusler, M. T., Decker, J. D. and Pfeffer, J. M. (2003). Social Learning for Collaborative Natural Resource Managemen. Society and Natural Resources 15: 309-326. Selfa, T. and Endter-Wada, J. (2008). The Politics of Community-Based Conservation in Natural Resource Management: A Focus for International Comparative Analysis. Environment and Planning A 40(4): 948-965. Sheppard, D. J., Moehrenschlager, A., McPherson, J. M. and Mason, J. J. (2010). Ten Years of Adaptive Community-Governed Conservation: Evaluating Biodiversity Protection and Poverty Alleviation in a West African Hippopotamus Reserve. Environmental Conservation 37(3): 270-282. Spencer, J. H. and Guzinsky, C. (2010). Periurbanization, Public Finance, and Local Governance of the Environment: Lessons from Small-Scale Water Suppliers in Gresik, Indonesia Environment and Planning A 42(9): 2131-2146. Spiteri, A. and Nepal, S. K. (2006). Incentive-Based Conservation Programs in Developing Countries: A Review of Some Key Issues and Suggestions for Improvements. Environmental Management 37(1): 1-14. Spiteri, A. and Nepal, S. K. (2008). Distributing Conservation Incentives in the Buffer Zone of Chitwan National Park, Nepal. Environmental Conservation 35(1): 76-86. Tai, H. S. (2007). Development through Conservation: An Institutional Analysis of Indigenous Community-Based Conservation in Taiwan. World Development 35(7): 11861203. Upreti, B. N. (1994). Royal Bardia National Park. Kathmandu, IUCN, Nepal. Webler, T., Kastenholz, H. and Renn, O. (1995). Public Participation in Impact Assessment: A Social Learning Perspective. Environmental Impact Assessment Review 15: 443-463. Wegge, P., Odden, M., Pokharel, C. P. and Storaas, T. (2009). Predator-Prey Relationships and Responses of Ungulates and Their Predators to the Establishment of Protected Areas: A Case Study of Tigers, Leopards and Their Prey in Bardia National Park, Nepal. Biological Conservation 142(1): 189-202. Wells, M. P. (1998). Institutions and Incentives for Biodiversity Conservation. Biodiversity and Conservation 7(6): 815-835. West, P., Igoe, J. and Brockington, D. (2006). Parks and Peoples: The Social Impact of Protected Areas. Annual Review of Anthropology 35: 251-277.

Complimentary Contributor Copy

In: National Parks Editor: Johnson B. Smith

ISBN: 978-1-62618-934-8 © 2013 Nova Science Publishers, Inc.

Chapter 4

PRIORITY AREAS INTEGRATION (PAI) METHOD: A TOOL TO FACILITATE BIODIVERSITY CONSERVATION? Georgios K. Vasios, Panayiotis G. Dimitrakopoulos† and Andreas Y. Troumbis‡ Biodiversity Conservation Laboratory, Department of Environment, University of the Aegean, Mytilini, Lesvos, Greece

ABSTRACT In biodiversity conservation planning a keystone process is the selection of the proper priority area and for that reason many diverse reserve selection (RS) methods have been developed. Usually depending on the chosen method a unique optimum solution is estimated based on the calculation of specific ecological criteria, indicators and algorithms. The objective of this chapter is the design of a method (called PAI) that integrates priority areas estimated from different RS methods. From this integration, through a simple process of computational steps, equivalent areas of priority are estimated which generate alternative conservation scenarios. The total area of conservation concern is separated into three major sub-priority areas (PA) called areas of high (H), medium (M) and low (L) priority. The high-PA includes all the sites that are always selected for conservation regardless of the RS method used, while the low-PA includes all the sites that are always rejected, respectively. Medium-PA includes all the sites that might be selected or rejected due to the RS method applied and their combination provides alternative solutions of priority areas. The insular areas of the Mediterranean basin were selected to implement the PAI method on a regional scale. In particular, in the South-East Aegean region a group of islands (Crete, Rhodes, Symi and Tilos) were studied as typical examples of combining 

Georgios K. Vasios. E-mail: [email protected]. Corresponding author: Panayiotis G. Dimitrakopoulos. E-mail: [email protected]; Tel.: + 30 22510 36236; Fax: + 30 22510 36263. ‡ Andreas Y. Troumbis. E-mail: [email protected]. †

Complimentary Contributor Copy

130

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis high values of biodiversity and strong economic development, mainly tourism. The results for all the islands show significant levels of concordance between potential sites. Spatial distribution of the various biodiversity values increases proportionally according to the island’s size, whereas in smaller islands the values from the different RS methods overlap. Alternative scenarios emerge for every potential size of the conservation area, providing the ability to implement socio-economic criteria into the process if needed, especially in areas undergoing intense development such as Crete and Rhodes. A generalized PA graph of the PAI method that analyses and presents a general spatial distribution of the conservation values of a conservation concern area is proposed. This unified graph, unique for each area of interest, visualizes the congruencies and variances of these values and based on their integration, produces conservation alternatives rather than defining an optimum solution. Such scenarios could facilitate the planners to design and implement more effective strategies in conservation planning.

Keywords: Conservation value, Crete, facilitation, quantitative methods, reserve selection, Rhodes, prioritization, biodiversity

1. INTRODUCTION Over the last decades, the most significant challenge in biodiversity conservation policymaking has been the definition of the boundaries and interrelations between scientific and non-scientific processes in species preservation and ecosystem services maintenance (Lawton, 1997; Margules and Pressey, 2000; Prendergast et al., 1999; Pressey and Bottrill, 2008; Reyers et al., 2010; Turnhout et al., 2007). Specifically the focus of modern conservation strategies has been on reducing biodiversity erosion stemming from changes in land-use and climate and affecting all levels of biological organization (Lindenmayer et al., 2010; Thuiller et al., 2008). Although future biodiversity loss and its distribution is difficult to assess without a high degree of uncertainty (May et al., 1995; Stork, 2010), the extinction rates of biodiversity will continue to increase with some estimations reporting that one-half of all species on earth may become extinct over the next 100 years (Lawton, 1997). In particular, where sufficient and systematic monitoring of species allocation exists, alarming trends have been recorded: e.g. the rapid decline of many European bird populations (Burfield and Van Bommel, 2004; Tucker and Heath, 1994) and the threat of extinction of approximately 15% of Mediterranean flora (Greuter, 1994; Lavergne et al., 2006). The actual conservation process for selecting a priority area is a problem involving the expert evaluation of the biodiversity value of the area and of consensus building among its users. Thus, this assessment depends on criteria i.e. nature reserve selection criteria, which despite appearing to be ‘scientifically objective’, are strongly subordinate to community perspectives and local stakeholder priorities (Lombard et al., 1997). Furthermore, the scientific value of these criteria is often debatable. Current scientific consensus argues that the conservation value assessment of an area should focus on regional biodiversity representativeness based on species richness quantification, and not on a species-specific basis (Arponen et al., 2010; Fleishman et al., 2006; Similä et al., 2006). Pressey et al. (1993) have developed the concept of representativeness by including the basic criteria of complementarity, flexibility and irreplaceability of the sites included into a regional network of nature reserves. Several scoring procedures for ranking areas of high value for biodiversity

Complimentary Contributor Copy

Priority Areas Integration (PAI) Method

131

conservation have been advocated, including (a) hotspots of richness: areas that contain the highest species richness (Curnutt et al., 1994; Lombard, 1995; Lombard et al., 1997; Margules et al., 2002; Medail and Quezel, 1997; Meliadou and Troumbis, 1997; Prendergast et al., 1993; Van Jaarsveld et al., 1998; Williams et al., 1996), (b) hotspots of rarity or endemism: areas that are richest in the most restricted distribution range species (Lombard, 1995; Lombard et al., 1997; Myers, 1990; Van Jaarsveld et al., 1998; Williams et al., 1996), (c) threatspots or hotspots of Red Data Book species: areas that include the highest richness of threatened species (Hernández-Manrique et al., 2012; Lombard, 1995; Troumbis and Dimitrakopoulos, 1998), and (d) areas of complementary richness: areas in which the maximum number of species is represented (Csuti et al., 1997; Freitag and Van Jaarsveld, 1997; Lombard, 1995; Lombard et al., 1997; Margules et al., 1988; Moilanen, 2008; Pressey and Nicholls, 1989; Pressey et al., 1997; Van Jaarsveld et al., 1998; Williams et al., 1996). Reserve selection (RS) algorithms are proposed to assist the identification and hierarchisation of potential conservation areas, in order to maximize the efficiency of conservation actions (Memtsas et al., 2002; Williams et al., 1996). The conservation objects used in these algorithms include species, vegetation types and landscapes. Furthermore, the design of conservation areas can be viewed as a multi-criteria decision problem that should combine ecological and sociopolitical criteria and a wide variety of multi-criteria methods devised over the last decades (Memtsas, 2003; Moffett and Sarkar, 2006). The choice of the appropriate technique, i.e. the method for the quantification of conservation value and the algorithm for the selection of conservation objects, depends on the size and resolution of data set, the ‘accepted’ level of representatitivity of the reserve system, and the time necessary to run the entire selection process (Geburek, 2010; Pressey et al., 1997). The problem of selection of a representative nature reserve system becomes more complicated in biodiversity-rich areas, which undergo rapid land-use change because of strong economic development policies. In these areas, it is important for land managers and environmental policy makers to be aware of that fact in order to better formulate their conservation goals and to correctly present to land users the real face of the conservation dilemmas (Oliver, 2002). The implementation of biodiversity conservation strategies to realworld actions needs to be supported by new tools and procedures that advocate the involvement of local communities in the decision making processes (Dimitrakopoulos et al., 2010; Knight et al., 2008; Oikonomou et al., 2011). Regardless of the hierarchization method used and the complexity introduced in the analysis, a specific ranking of the potential sites for conservation is always created. The final area is selected based on such ranking of sites for a specific size of the conservation concern area. The goal in every RS method is to separate the area of interest in two distinct sub-areas: the Conserved area (C) and the Rejected area (R) (Figure 1a). But how common and distinct are the sites selected by the various RS methods? Orme et al. (2005) used Venn diagrams and congruence graphs to visualize congruence between multiple RS methods; Dimitrakopoulos et al. (2004) employed overlap estimation and Spearman correlation coefficients to calculate variances. Additionally, Abellan et al. (2005) used cumulative species graphs to visualize the result of a specific RS method on biodiversity conservation demonstrating the final impact of the conservation planning process. Each priority area as a result of a specific RS method calculates different aspects of the biodiversity value of the conservation concern area and a method that could compare the results from multiple RS methods could facilitate the conservation planning process.

Complimentary Contributor Copy

132

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

The aim of this chapter is to develop a new method, called Priority Areas Integration (PAI) method that (a) facilitates conservation planning by integrating various biodiversity values of a conservation concern area; (b) analyzes and visualizes the site hierarchization process of multiple RS methods for every potential size of the conservation area and (c) creates alternative conservation scenarios based on the integration of these biodiversity values. The basic steps and graphs of the PAI method are analyzed and presented using an explanatory hypothetical case study in Section 3. A cluster of four islands (Crete, Rhodes, Symi and Tilos) in the South East Aegean region was selected to implement the PAI method on a regional scale as typical example of the insular areas of the Mediterranean basin with extremely high and threatened plant diversity combined with strong economic development, mainly tourism (Medail and Quezel, 1997). Finally, a generalized PA graph of the PAI method that analyses and presents a general spatial distribution of the biodiversity values of a conservation area is proposed in Section 4.

2. METHODS 2.1. General Concept of Priority Areas Integration (PAI) Method Let us assume that a conservation concern area is divided into N number of sites of equal size. The use of multiple reserve selection (RS) methods will give different grades for these sites creating multiple set of site ranking. The goal of the Priority Areas Integration (PAI) method is to create alternative solutions of conservation by comparing these multiple RS methods that convey the various biodiversity values of a conservation area. In detail, the PAI method separates the conservation concern area into three major sub-areas, called priority areas (PA) of high, medium and low priority. The High-PA (H) includes the sites that are always selected for conservation regardless of the RS method used and the Low-PA (L) includes all the sites that are always rejected. Medium-PA (M) is all the sites that might be selected or rejected based on the RS method applied and their combination creates alternative solutions for conservation (Figure 1b). For every alternative, the Medium-PA leads to the formation of the Selected-PA and Gap-PA. The Selected-PA (S) is the Medium-PA sites that are selected based on the applied RS method and the Gap-PA (G) is the Medium-PA sites that are rejected but were potential candidates for conservation if other scenarios would have been chosen. Through this process and for each alternative, the conservation concern area is divided into the final Conserved areas (C) and Rejected areas (R). The first is divided into the H and S priority areas and the second into the L and G priority areas for the same alternative.

2.2. Priority Areas (PA) Graph The PAI method analyses the conservation decision process for a specific set of RS methods with the use of the PA graph. For every potential size of the conservation area, the PA graph visualizes the allocation of all sites to the various PA. In detail, the PA graph is formed by a square with the use of a Cartesian coordinate system.

Complimentary Contributor Copy

Priority Areas Integration (PAI) Method

133

Figure 1. Segregation of the conservation concern area by: (a) specific method into the Conserved areas and Rejected areas; (b) PAI method into the various types of priority areas (High-PA, Medium-PA, Low-PA, Selected-PA and Gap-PA) (see Table 1 for abbreviations).

In the x-axis is represented the size of the selected conservation area as a percentage of the total concern area for conservation with values from 0% to 100%. In the y-axis is represented the cumulative size of the various types of PA (H, M and L) that divide the total area of interest taking values from 0 to N number of sites. For every size of the final conservation area (x-axis), the size of the priority areas (y-axis) varies. If all these solutions are calculated and displayed on the PA graph three curves are formed: (a) the Conserved Areas curve (CA); (b) the Areas Congruence curve (AC) and (c) the Areas Variance curve (AV). The form of a general PA graph is presented in Figure 2 and described in details in Section 4.3 with the curve functions and main abbreviations used presented in Table 1. These three curves separate the square of the PA graph into various sub surfaces -one for each priority area- that visualize the modification of each PA due to variations in the conservation area size underscoring the interdependence between all PA. The PA graph can also be viewed as a summation of N columns where each column is a stack of all the N sites of the total conservation concern area. The NxN sites form the square of the PA graph where the horizontal side coincides with the x-axis and the vertical side with y-axis of the PA graph. For a specific conservation area size, a unique column of N sites is created from all the sites of the total area by classifying every site to a specific PA based on the comparison of the RS methods. As the conservation size changes, the various PA are modified in size and shape due to the new distributions of site classification into the different PA. Examples of PA graphs as stack of sites are presented in Sections 3.4 and 4.1.

Complimentary Contributor Copy

134

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

Figure 2. The PAI method (for k RS methods) of a conservation concern area (N sites). Four PA graphs presenting the equivalent basic phases of the PAI method: (a) Conserved Areas; (b) Areas Congruence; (c) Areas Variance and (d) Priority Alternatives (see Table 1 for abbreviations).

Complimentary Contributor Copy

Priority Areas Integration (PAI) Method

135

Table 1. List of the basic terms used in the PAI method with their abbreviations, descriptions and functions Abbreviation Description Function and Equation General PA Priority Area PAI Priority Areas Integration RS Reserve Selection Type of areas with conservation interest (number of sites or % of area) C Conservation area C = H+S R Rejected area R = L+G H High Priority Area (High-PA) M Medium Priority Area (Medium-PA) M = S+G L Low Priority Area (Low-PA) S Selected Priority Area (Selected-PA) S = M-G G Gap Priority Area (Gap-PA) G = M-S Variables of PAI method minR minimum Ranking value of a site from k methods minR = min{R1, R2, …, Rk} maxR maximum Ranking value of a site from k methods maxR = max{R1, R2, …, Rk} a, b, …, o name of sites used in the explanatory data set A Alternative selection of sites (number of scenarios) A = C(M, S) = M! / (S!*G!) C0 specific Conservation area (% of total area) N0 specific conservation area (Number of sites) NH High-PA (Number of sites) NH = H NM sum of High-PA and Medium-PA (Number of sites) NM = H+M N total area of conservation interest (Number of sites) N = C+R, N = H+M+L M' max. Medium-PA (number of sites) M' = S'+G' S' max. Selected-PA of M' (number of sites) G' max. Gap-PA of M' (number of sites) C'0 Conservation area of M' (% of total area) Smax max. Selected-PA (% of conservation area) Gmax max. Gap-PA (% of conservation area) Curves of PAI graph CA Conservation Areas curve CA: N0 = f(C0), C0є[0, 100] AC Areas Congruence curve AC: NH = f(C0), C0є[0, 100] AV Areas Variance curve AV: NM = f(C0), C0є[0, 100]

2.3. Basic Phases of PAI Method The PAI method is completed in four phases which are (1) the Conserved Areas phase; (2) the Areas Congruence phase; (3) the Areas Variance phase and (4) the Priority Alternatives phase. For a set of RS methods and a specific size of conservation area these four phases can be described as the process of answering four distinct questions respectively. These questions are: (1) which area is proposed for conservation from each RS method?; (2) do these potential solutions concur?; (3) do these potential solutions vary? and (4) what

Complimentary Contributor Copy

136

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

are the total alternative conservation areas that could be proposed? These phases separate the total area into distinct PA only for a specific size of conservation area. If the size of this area is changed the application of the four phases of PAI method might give a new set of PA. For a conservation concern area consisted of N sites, the process must be repeated N times respectively for calculating all possible PA. To avoid such complexity in the analysis, a simplified process has been designed to calculate the PA for all potential sizes of conservation area. These calculations are based on the introduction of the three PAI curves which are calculated with the use of two new indices, called High-PA and Medium-PA indices. The basic steps and graphs to compute the four phases of PAI method, the two PA indices and the three PAI curves are analyzed in detail using an explanatory hypothetical case study presented in Section 3.

2.4. Case Study For the implementation of PAI method on a regional scale for facilitating a real conservation planning decision process, a cluster of four islands (Crete, Rhodes, Symi and Tilos) in the South East Aegean region was selected. These islands are typical examples of the insular areas of the Mediterranean basin and they are characterized by extremely high and threatened plant diversity (Medail and Quezel, 1997). The biodiversity values such as species richness, distribution and vulnerability of the conservation concern areas are rated using the three most representative RS indices which are (1) the species richness index; (2) the range-size rarity index and (3) the vulnerability index (Dimitrakopoulos et al. 2004). Furthermore, on all four islands the natural environment is under significant pressure from economic development, mainly tourism but also from agriculture, urban expansion and forest fires (Dimitrakopoulos et al., 2004, 2005). Such a strong human presence in these areas raises barriers to a successful conservation process and new alternatives could facilitate the improvement of the conservation planning of these islands. Zero-one matrices and ranking of species richness, range-size rarity and vulnerability for all four islands had been calculated in previous studies by Dimitrakopoulos et al. (2004; 2005). The main island of Crete with the surrounding small islands and islets were divided into 162 sites of 8.25 km2 (2.87 x 2.87 km) each and its vascular flora comprises 1624 native species and 76 species introduced by man within an area of 8700 km2 (Dimitrakopoulos et al., 2004; Turland et al., 1993). The main island of Rhodes with the surrounding islets were divided into 118 sites of 16 km2 (4 x 4 km) each and its vascular flora comprises 1120 species, native and introduced by man, within an area of 1400 km2 (Carlstrom, n.d.; Dimitrakopoulos et al., 2005). Additionally, the islands of Tilos and Symi with the surrounding islets are divided into 6 and 13 sites of 16 km2 each and their vascular flora comprises 369 and 415 species within an area of 63 km2 and 58 km2 respectively. These results were used as the initial data set for the case study analysis. The PAI method process was implemented in a spreadsheet.

Complimentary Contributor Copy

137

Priority Areas Integration (PAI) Method

3. CALCULATIONS The PAI method with its basic steps and graphs are analyzed and presented with an explanatory hypothetical data set. The main objective is to describe in detail: (a) the formation of the various PA (H, M and L) for a specific size of the conservation area (one column in the PA graph); (b) the introduction of two new indices, called High-PA and Medium-PA indices; and (c) the calculation and the formation of the various PA (H, M, L) and the PAI curves (CA, AC, AV) for every potential size of the conservation area (N columns in the PA graph). In our explanatory case study, the hypothetical area has been divided into 15 sites (N = 15) of equal size (grid cells) labeled: a, b, c, through to m, n and o (Figure 3a). In each site, we assume that the species presence had been recorded creating a zero-one matrix and that with the use of three different hypothetical RS methods, the 15 sites were graded and ranked in three separate ways.

3.1. Four Phases of PAI Method for a Specific Size of the Conservation Area The PAI method is completed in four phases which are (1) the Conserved Areas phase; (2) the Areas Congruence phase; (3) the Areas Variance phase and (4) the Priority Alternatives phase. For a specific size of conservation area and with the use of our explanatory data set, the rearrangement of sites that occurs in each phase is displayed in Table 2 and analyzed in detail by answering four distinct questions, one for each phase, as described in Section 2.3.

1st – Conserved Areas Phase If the goal is to conserve a specific conservation area C0 (N0 sites) of the total area (N sites), for example, 7 out of 15 sites (C0 = 47%) then based on three hypothetical RS methods used in our case study, there are three different solutions of 7 sites that could be conserved in each case (Table 2: 1st phase).

a

b

Figure 3. The basic characteristics of the hypothetical area in the explanatory data set: (a) map of the area divided into 15 sites of equal size (grid cells) labeled: a, b, c, through to m, n and o; (b) sites rearranged into three distinct PA: High-PA (colored black surface), Medium-PA (gray) and Low-PA (white) for a specific conservation area (C0 = 47%) (see Table 1 for abbreviations).

Complimentary Contributor Copy

138

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

Regardless of the RS method used, the total area will be separated into the Conserved areas (C) and Rejected areas (R). Even though they may contain different sites, these subareas have the same sum of sites where C area is equal to 7 sites and R area to 8. Although each of these three Conserved areas offer the N0 sites with the highest biodiversity value, only one area can be selected which will be based on the RS method assumed to be the most suitable to implement specific conservation strategy.

2nd – Areas Congruence Phase Do these three Conserved areas selected by the three different RS methods concur in any way? In the 1st phase (Table 2), we notice that 4 sites (labeled: b, m, l, c) are common in all three RS methods. These 4 sites will be called the High-PA (H) as they are always selected for conservation regardless of the RS method. This distinction separates each Conserved area to the High-PA (4 sites) and the Selected-PA (S) (3 sites) by rearranging the ranking order of the 7 sites of the Conserved area created from each RS method respectively (Table 2: 2nd phase). Because the sites of the High-PA are identical for all the RS methods used they are grouped together in one column for simplification purposes, unlike the sites of the SelectedPA that contain three different set of sites for every RS method. 3rd – Areas Variance Phase Do these three Conserved areas vary in some way? Focusing on the Selected-PA of the 2nd phase (Table 2), there are three set of 3 sites {(j, e, f), (g, j, i), (e, i, g)}, one for each RS method. Even though each set of sites is unique, all of them could be created from equivalent combinations of 3 sites from a set of 5 (g, j, e, i, f). This set of 5 sites will be called the Medium-PA (M) i.e. the sites that are potential candidates for selection. Based on a specific RS method, some of the sites of the Medium-PA will be selected and the rest will be rejected. The sites of the Medium-PA are rearranged as a group after the High-PA sites and before the remaining sites of the total area from top to bottom in the 3rd phase (Table 2). The remaining sites, 6 in number in this hypothetical case study will be called the Low-PA (L) and they are the sites that are always rejected regardless of the RS method used. Because the sites of the Low-PA are identical for all the RS methods used they are grouped together in one column for simplification purposes too. The rearrangement of sites separates the total area into three distinct PA of high, medium and low priority (Figure 3b) and forms the final order of the 15 sites column for a specific conservation area (47% of total) as shown in the PA graph (Figure 4). 4th – Priority Alternatives Phase Do any new alternative solutions arise from the segregation of the total area into three distinct PA (H, M and L)? In the Medium-PA of the 3rd phase (Table 2), we notice that 10 unique unordered combinations of 3 sites can be created from the specific set of 5 sites (details of equation in Table 1). These combinations are all the potential alternative solutions including the three initial solutions generated from the RS methods for the specific conservation area (Table 2: 4th phase). When a specific alternative is chosen, the Medium-PA is separated into a Selected-PA (3 sites) and a Gap-PA (G) (2 sites).

Complimentary Contributor Copy

Table 2. The four phases of the PAI method for a specific size of the conservation area (C0 = 47%) of the explanatory hypothetical data set. The process of the sites rearrangement is presented in each phase as the various types of priority areas (H, M, L, S and G) are formed step by step (see Table 1 for abbreviations) 1st Phase Conserved Areas Methods Areas Ranking st 1 2nd 3rd 1 b m b 2 m b m 3 c g c Conserved 4 j l e (C) 5 l j l 6 e c i 7 f i g 8 g k j 9 h e f 10 k h k d d h Rejected 11 (R) 12 i f a 13 a a n 14 n n d 15 o o o

2nd Phase Areas Congruence Methods Areas 1st 2nd 3rd b m High-PA (H) l c j g e Selected-PA e j i (S) f i g g k j h e f k h k d d h Rejected (R) i f a a a n n n d o o o

3rd Phase Areas Variance Methods Areas Integration b m High-PA (H) l c g j Medium-PA e (M) i f k h a Low-PA (L) d n o

4th Phase Priority Alternatives Alternatives (A) Areas 1 2 3 4 5 6 7 8 b m High-PA (H) l c g g g g g i f g Selected-PA j j j i f j j i (S) e i f e e e e f i e i j i g i j Gap-PA (G) f f e f j f g e k h a Low-PA (L) d n o

Complimentary Contributor Copy

9

10

i j f g e

i f e g j

140

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

3.2. High-PA and Medium-PA Indices In our hypothetical area, each site takes a particular grade based on its ranking position for each RS method (Table 3). The lowest grade of a site signifies the best ranking position. For example, when the 1st and 3rd RS methods are used, the grade of site b is equal to one (best grade) because it will be the first to be selected for conservation. When the 2nd RS method is used its grade equals two, meaning that site b will be the second choice after site m. These ranking grades of all sites (Table 3) are equivalent to the ranking of sites in the 1st phase of Table 2, viewed in reverse. In Table 3, sites are ordered based on their labels and followed by their ranking grades from each RS method, instead of Table 2 where the sites are ordered based on their ranking grade (from best -grade number 1- to worst -grade number 15). An interesting attribute of the ranking grade of a site is that it determines the size (N0) that the conservation area must have so that this site will be selected for conservation. Comparing the ranking grades of a specific site we can discern from which method the site gets its highest (maxR) or lowest (minR) ranking grade (Table 3). For example, site g gets its best ranking grade (minR = 3) when the 2nd RS method is used and its worst grade (maxR = 8) when the 1st RS method is used, respectively. When the size of the conservation (C0) area is less than 3 sites (N0 < 3) then site g is rejected and will be part of the Low-PA. When C0 area contains 3 sites (N0 = 3) then site g will be selected for conservation and will be part of the Medium-PA due to 2nd RS method. When C0 area increases to 7 sites (N0 = 7) then site g will be selected by the 2nd and 3rd RS method and when C0 area expands to 8 sites (N0 = 8) then site g will be selected by all the RS methods. This means that site g will be part of the High-PA. The highest ranking grade (maxR) of a site determines the size (N0) that the conservation area must have so that the site will be selected for conservation regardless of the RS method used. When the size of the conservation area (N0) is equal to the maxR of a site (N0 = maxR) then this site will be part of the High-PA. This attribute of the maxR is used to calculate the High-PA for all potential sizes of the conservation area. A new index called ‘High-PA index’ is calculated by ranking all sites based on their maxR in ascending order (from low to high values) as presented in Table 4. For example, site g (with maxR equals 8) is ranked in 5th place in the High-PA index which means that it will be the 5th site to be included into High-PA when conservation area is equal to 53% of the total area (i.e. N0 = 8 and C0 = 53% in Table 5). Respectively, the lowest ranking grade (minR) of a site determines the size (N0) that the conservation area must have so that the site will be selected by at least one RS method. When the size of the conservation area (N0) is equal to the minR of a site (N0 = minR) then this site will be part of the Medium-PA. This attribute of the minR is used to calculate the MediumPA for all potential conservation areas. A new index called ‘Medium-PA index’ is calculated by ranking all sites based on their minR in ascending order (from low to high values) as presented in Table 4. For example, site g (with minR equals 3) is ranked in 4th place in the Medium-PA index which means that it will be the 4th site to be included into Medium-PA when conservation area is equal to 20% of the total area (i.e. N0 = 3 and C0 = 20% in Table 5).

Complimentary Contributor Copy

141

Priority Areas Integration (PAI) Method

Table 3. The ranking grades of sites produced by three hypothetical RS methods used in the explanatory hypothetical data set and the calculated maxR and minR values of each site which are used for the calculation of the High-PA and Medium-PA indices (see Table 1 for abbreviations)

Sites a b c d e f g h i j k l m n o

1st 13 1 3 11 6 7 8 9 12 4 10 5 2 14 15

Methods 2nd 13 2 6 11 9 12 3 10 7 5 8 4 1 14 15

3rd 12 1 3 14 4 9 7 11 6 8 10 5 2 13 15

Ranking Grades Maximum (maxR) 13 2 6 14 9 12 8 11 12 8 10 5 2 14 15

Minimum (minR) 12 1 3 11 4 7 3 9 6 4 8 4 1 13 15

Table 4. The calculation of the High-PA and Medium-PA indices for the explanatory data set. The sites are ranked based on their maxR value for High-PA index and minR for Medium-PA index respectively (see Table 1 for abbreviations) Indices of PA Ranking 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

maxR 2 2 5 6 8 8 9 10 11 12 12 13 14 14 15

High-PA Sites b m l c g j e k h i f a d n o

minR 1 1 3 3 4 4 4 6 7 8 9 11 12 13 15

Medium-PA Sites b m c g l j e i f k h d a n o

Complimentary Contributor Copy

142

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

3.3. Calculation of the PA, PAI Curves and Alternatives With the use of the High-PA and Medium-PA indices (Table 4), the parameters of the PA graph are calculated for each size of the conservation area (N0) (Table 5): (1) the size of the High-PA (H) with their new sites; (2) the size of the Medium-PA (M) with their total sites; (3) the values N0, NH, and NM of the PAI curves CA, AC and AV; (4) the sizes of the Selected-PA (S) and Gap-PA (G) and (5) the number of alternative (A) solutions. For the calculation of High-PA and Medium-PA in Table 5, two mathematical rules are used in the data set of the High-PA and Medium-PA indices in Table 4: a b

High-PA: for a specific N0, the High-PA contains only the sites that have maxR less than or equal to N0 (maxR ≤ N0). Medium-PA: for a specific N0, the Medium-PA contains only the sites that have minR less than or equal to N0 (minR ≤ N0) and maxR higher than N0 (maxR > N0). The last condition excludes from the Medium-PA all the sites that are already included in the High-PA for the specific conservation area (N0).

For example, for a conservation area equal to 47% of the total area (i.e. N0 = 7 and C0 = 47% in Table 5), the sizes of the High-PA and Medium-PA are calculated as followed: a

b

High-PA: when N0 = 7 then maxR ≤ 7. Based on Table 4, only the sites with maxR ≤ 7 are included to the High-PA which creates a set of 4 sites containing sites b, m, l and c. Medium-PA: when N0 = 7 then minR ≤ 7 and maxR > 7. Based on Table 4, only the sites with minR ≤ 7 are included to the Medium-PA, excluding the sites with maxR ≤ 7 (sites b, m, l and c). These conditions create a set of five sites containing sites g, j, e, i and f.

For the same conservation area (N0), based on the calculations of the High-PA and Medium-PA and the use of the functions of Table 1, can easily be calculated (Table 5): a b c d e f g

the value of the CA curve which equals to N0 = 7 the value of the AC curve which equals to NH = 4 the value of the AV curve which equals to NM = H + M = 9 the size of the Low-PA (L = N – NM = 6) which contains sites k, h, d, a, n and o the size of the Selected-PA (S = N0 – NH = 3) which contains 3 sites the size of the Gap-PA (G = M – S = 2) which contains 2 sites the sum of all the alternative solutions (A = 10) that can be produced from the set of 3 sites out of 5 which creates ten unique set of 3 sites (Selected-PA)

3.4. Construction of Priority Areas (PA) Graph In our explanatory case study, for every potential size of conservation area, the total area of the 15 sites is divided into three sub-areas of high, medium and low priority based on the integration of three hypothetical RS methods (presented in Section 3.3).

Complimentary Contributor Copy

143

Priority Areas Integration (PAI) Method

Table 5. The calculation of the PA graph parameters (N0, C0, H, new High-PA sites, M, NM, total Medium-PA sites, S, G, L and A which is the number of alternatives) using the High-PA and Medium-PA indices for the explanatory hypothetical data set (see Table 1 for abbreviations) Conservation Area C0 N0 (%) 1 7 2 13 3 20 4 27 5 33 6 40 7 47 8 53 9 60 10 67 11 73 12 80 13 87 14 93 15 100

High-PA H New (NH) Sites 0 2 b, m 2 2 3 l 4 c 4 6 g, j 7 e 8 k 9 h 11 i, f 12 a 14 d, n 15 o

Priority Areas Medium-PA

Low-PA

Alternatives

M

NM

Total Sites

S

G

L

A

2 0 2 5 4 4 5 4 4 3 3 2 2 0 0

2 2 4 7 7 8 9 10 11 11 12 13 14 14 15

b, m

1 0 1 2 2 2 3 2 2 2 2 1 1 0 0

1 0 1 3 2 2 2 2 2 1 1 1 1 0 0

13 13 11 8 8 7 6 5 4 4 3 2 1 1 0

2 1 2 10 6 6 10 6 6 3 3 2 2 1 1

c, g c, g, l, j, e c, g, j, e g, j, e, i g, j, e, i, f e, i, f, k i, f, k, h i, f, h i, f, d d, a d, n

In Figure 4, the square of the PA graph can be formed based on the calculations in Table 5, as a summation of N columns where each column is a stack of all the N sites of the total conservation concern area. In the y-axis of Figure 4, the cumulative sizes of the various PA (H, M and L) are presented forming the PAI curves (three lines in the PA graph). By increasing C0 (x-axis), new sites of the High-PA are introduced, forming the High-PA surface (H) in the PA square (bottom-right corner). The sites of the Medium-PA are also calculated, forming the Medium-PA surface (M) in the PA square (diagonal area). The remaining area of the PA square (top-left corner) forms the Low-PA surface (L) which contains the sites of the Low-PA. An example of a complete column of sites rearrangement on the various PA (H, M and L) is visualized in the PA graph (Figure 4) for C0 = 47% (N0 = 7) based on the Table 2 data set. In Figure 4, all the Medium-PA sites for every potential size of the conservation area are shown. For the High-PA, only the new High-PA sites are shown for simplification reasons. For the same purpose, none of the Low-PA sites are displayed except for the example of a complete column of sites when C0 = 47%. The High-PA (the surface in the bottom-right corner) covers the biggest part of the Conserved area indicative of the high rates of congruence between the hypothetical RS methods. In three cases of C0 (equal to 13%, 93% and 100%) the High-PA coincides with the Conserved area without creating any Medium-PA (M = 0 in Table 5). In two cases (C0 equals to 27% and 47%) the Medium-PA takes its highest value (M = 5 in Table 5) producing the maximum number of alternative solutions (A = 10) but only in one of these cases (C0 = 47%) the Selected-PA gets its highest value (S = 3). For this specific size of the conservation (C’0)

Complimentary Contributor Copy

144

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

area, the maximum Medium-PA (M’ = 5) and the maximum Selected-PA (S’ = 3) are formed which means that 3 sites out of 7 (71% of the Conserved area) varies highlighting the conservation area size with the maximum variation between the RS methods.

4. RESULTS AND DISCUSSION The four islands of our case study (Crete, Rhodes, Symi and Tilos) are analyzed with the use of the PAI method by integrating three RS methods which are the (1) species richness index; (2) range-size rarity index and (3) vulnerability index. This process for each island created distinct PA graphs. The results are grouped and compared in pairs based on the islands’ size similarity (Tilos with Symi and Crete with Rhodes).

4.1. Tilos and Symi In Tilos and Symi, the total conservation concern area (N) is divided into a different number of sites but with identical site size per island (same grid of cells). Therefore, on Tilos, N is equal to 6 sites and on Symi to 13 sites. For every C0, all the sites are rearranged to various PA dividing the PA square into the High-PA, Medium-PA and Low-PA.

Figure 4. The PA graph of the explanatory data set presenting the sites rearrangement to the various types of PA (number of sites) for every size of the conservation area (% of total area). The total Medium-PA sites for each conservation area are shown. For simplification reasons, only the new HighPA sites are shown and none of the Low-PA sites, except of a complete column of sites rearranged into the three distinct PA: High-PA (colored black surface), Medium-PA (gray) and Low-PA (white) when conservation area equals to 47% (C0 = 47%) (see Table 1 for abbreviations).

Complimentary Contributor Copy

145

Priority Areas Integration (PAI) Method

In Figure 5, two PA squares are formed, one for each island where the sites’ rearrangement into the High-PA and Medium-PA for all potential C0 and the CA curve of the Conserved sites (diagonal line) are illustrated. The sites’ rearrangement into the Low-PA is hidden even though it is part of the PA square for simplification reasons. In Tilos, for all three RS methods the ranking of the sites is identical enhancing the high correlation of the biodiversity values for each site. Therefore, in the PA graph only the HighPA and Low-PA are formed which determines the final Conserved and Rejected areas. In Symi, for all three RS methods the ranking of the sites is identical for most of the conservation areas (8 out of 13 potential sizes), indicating significant correlation of the biodiversity values for each site. In the PA graph (Figure 5b), the Medium-PA is formed in two area groups, one from 3 to 5 potential conservation sites and the second from 8 to 9 sites. In detail, when C0 is equal to 23% and 62%, the maximum Medium-PA (M’) (3 sites) and the maximum Gap-PA (G’) (2 sites) are formed. In these cases, the highest variation between the three RS methods (1 site for each RS method) emerges. The only 1 site of the Selected-PA (S) enhances the congruence between the three RS methods used.

a

b

a

Tilos. b Symi. Figure 5. The PA graphs of the Mediterranean islands of: (a) Tilos with a total area of 6 sites and (b) Symi with 13 sites. For each potential conservation area, all sites of the High-PA and Medium-PA are presented except of the Low-PA sites which are hidden to simplify the graph (see Table 1 for abbreviations).

Complimentary Contributor Copy

146

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

4.2. Comparison of Crete and Rhodes In Crete and Rhodes, the total conservation concern area is divided into a different number of sites with diverse site size per island. Therefore, N is equal to 162 sites on Crete’s PA graph and 118 sites on Rhodes’ PA graph. For every C0, all the sites are rearranged to various PA dividing the PA square into the High-PA, Medium-PA and Low-PA (Figure 6a). The size of the Medium-PA changes in both islands as the C0 increases with a variation from 0 to 24 sites for Crete and from 0 to 13 sites for Rhodes. When the Medium-PA doesn’t exist (M = 0) then for that specific size of C0 all three RS methods select the same set of sites for conservation, creating the same High-PA. This phenomenon is rare for the specific islands and takes place in Crete only for two different sizes of C0 (0.6% and 1.2% of the total area) and in Rhodes for five (0.8%, 3.4%, 4.2%, 94.9% and 99.2% of the total area). In Crete, the maximum Medium-PA (M’) is equal to 24 sites and the maximum Selected-PA (S’) is equal to 12 sites (13.2% of the Conserved area) when the conservation area (C’0) is equal to 56.2% of the total area. Respectively in Rhodes, the maximum Medium-PA (M’) is equal to 13 sites and the maximum Selected-PA (S’) is equal to 6 sites (7.9% of the Conserved area) when the conservation area (C’0) is equal to 64.4% of the total area. The Gap-PA (G) and Selected-PA (S) have been visualized as a percentage of the Conserved area (Figure 6b; 6c) enhancing the High-PA, Selected-PA and Gap-PA variation for each modification of the size of the conservation area. With these graphs, the congruencies and variances between the PA for every conservation area size are more clearly illustrated facilitating the overall analysis of the conservation process (Orme et al., 2005). The Gap-PA varies from zero to 75% and 50% of the total area of Crete and Rhodes respectively. The highest value of Gap-PA (Gmax) is present when the conservation area (C0) is equal to 4.9% and 1.7% of the total area of Crete and Rhodes respectively (Figure 6b). In Figure 6(c), the Selected-PA varies from zero to 69% and 50% of the total area of Crete and Rhodes respectively, having the highest value of Selected-PA (Smax) when C0 is equal to 8.0% and 1.7% of the total area of Crete and Rhodes respectively. In both islands, the High-PA is larger than the Selected-PA for most of the conservation areas as illustrated in all the graphs of Figure 6. Additionally, there is a strong symmetry between the Gap-PA and Selected-PA (Figure 6b; 6c). These results of the PA shows the high level of congruence between the three RS methods used for all the sites of Crete and Rhodes. Such a correlation was highlighted by Dimitrakopoulos et al. (2004; 2005) in earlier studies. The Medium-PA of Crete is larger than Rhodes which signifies that biodiversity values of sites on Crete compared to Rhodes have a higher rate of variation between them for most of the potential sizes of conservation area. These results are shown even more clearly on the Selected-PA graphs (Figure 6c) by comparing the size of High-PA and Selected-PA of the two islands, especially for low size conservation areas. Furthermore, the degree of change for the Gap-PA and Selected-PA is different for the two islands, as the size of the conservation area increases (Figure 6b; 6c). In Crete, these PA have decrease steadily from the highest value (Gmax and Smax) to zero value, as the C0 increases. In Rhodes, as the C0 starts to increase there is a strong variation on the rate of change for the percentages of the Gap-PA and Selected-PA, moving from 0.8% to 21.2% of the Conserved area. As the C0 continues to increase, the Gap-PA and Selected-PA decrease at a similar rate as in Crete. This asymmetry between the two islands indicates that in the case of Crete, the differences in the sites selected by the three RS methods decrease, as the size of the

Complimentary Contributor Copy

Priority Areas Integration (PAI) Method

147

conservation area is increased. For Rhodes, when the conservation area equals 3.4%, 4.2% and 19.5% then the RS methods used give a similar set of Conserved sites. These conservation areas combine high level of biodiversity value based on all three RS methods used with low levels of area size.

Figure 6. The various PA graphs of the case study of the Mediterranean islands Crete and Rhodes: (a) PA squares presenting the various PA (number of sites), (b) Gap-PA graph (% of Conserved area) and (c) Selected-PA graph (% of Conserved area) of Crete (left column of graphs) and Rhodes (right column) (see Table 1 for abbreviations).

Complimentary Contributor Copy

148

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

4.3. PAI Method General PA Graphs The PAI method analysis for the four islands of our case study creates four unique sets of PA graphs, one set for each island. These graphs are formed by the implementation of selected conditions i.e. specific RS methods and grid-cells size, to local biodiversity characteristics. Based on these results we could create a generalized set of PA graphs for a theoretical case study. Let us assume that there is a hypothetical conservation concern area that is divided into N sites of equal size with a zero-one matrix of species presence constructed. All the sites of this area are graded and ranked k times with the use of k number of RS methods, respectively. Combining the four steps of the PAI method analysis for a specific conservation area C 0, as presented in Section 3 (Table 2), for all the potential sizes of the conservation area (Table 5) with the use of the PA graph, four unique PA graphs are produced. These PA graphs, one for each phase of the PAI process, visualize the theoretical general version of the PAI method for a hypothetical conservation concern area (Figure 2).

1st – Conserved Areas Phase If the goal is to conserve a specific conservation area C0 (N0 sites) of the total area (N sites) then regardless of the RS method used the total area will be separated into a Conserved area (C) and a Rejected area (R). Even though they may contain different sites, these subareas have the same sum of sites where area C is equal to N0 sites and area R to N-N0 sites. For all potential C0 areas, the N0 sites are calculated for every RS method separately and can be viewed in a unified PA graph with equivalent CA curves respectively (Figure 2a). In the PA square, the shape of these CA curves is a diagonal line (N0 = C0). The result from all the calculations is a set of k diagonal lines, one for every RS method that overlap with each other in the unified PA square. In detail, due to the attribute of the PA square as a summation of N columns of N sites, for the k RS methods a set of k unique PA graphs are formed with a different rearrangement of sites, even though they are represented with the same diagonal line. 2nd – Areas Congruence Phase Do all the potential Conserved areas from the k RS methods used concur in any way? By calculating the number of sites of High-PA (NH) for all potential C0 areas, a unique AC curve is designed (Figure 2b). The High-PA (H) that is formed in all the k PA graphs is unique having the same shape and containing the same rearrangement of sites for all k RS methods. However, the Selected-PA (S) which is formed between the CA and AC curve, even though it has the same size in all the k PA graphs, contains a different set of sites for every RS method. 3rd – Areas Variance Phase Do all the potential Conserved areas from the k RS methods vary in some way? By calculating the number of sites of High-PA and Medium-PA (NM) for all potential C0 areas, a unique AV curve is designed (Figure 2c). The Medium-PA (M) that is formed in all the k PA graphs is unique having the same shape and containing the same set of sites for all k RS methods. Respectively, the Low-PA (L) that is formed is unique and the same (in shape and set of sites) for all k RS methods.

Complimentary Contributor Copy

Priority Areas Integration (PAI) Method

149

4th – Priority Alternatives Phase Do any new alternative solutions arise from the analysis of the total area into the three distinct PA (H, M and L)? For a specific conservation area from all the potential combinations of sites of the Medium-PA, alternative solutions are generated including the initial solutions from the k RS methods. When a specific alternative is chosen the MediumPA is separated into a Selected-PA (S) (with N0-NH sites) and a Gap-PA (G) (with NM-N0 sites) (Figure 2d).

4.4. Conservation Planning Facilitation The PAI method analyses and presents the spatial distribution of the biodiversity values of a conservation concern area. For all the potential sizes of the selected conservation area, a unique set of PA graphs of the total area is produced as a combination of (a) specific species that have been identified and spatially recorded; (b) the set of the RS methods used for analysis and (c) the size of the grid-cell proportionate to the size of the total conservation concern area. If any of these features are modified, this will have an effect on the shape of the curves of PA graphs and respectively on the size and spatial distribution of High-PA, Medium-PA and Low-PA. Which features are more critical to the modification of the various PA? A sensitivity analysis of these parameters will enhance such variations and could become a subject of future research. The presence of key species that change the ranking of sites calculated by the RS methods used due to their biodiversity attributes (e.g. rarity, vulnerability) might have a significant effect on the final classification of sites to the various PA (Blasi et al., 2011; Margules et al., 2002). The RS methods, which determine the size and spatial distribution of the various priorities areas, are selected to implement specific conservation goals (Moilanen, 2008). Our analysis focused on the use of the most ‘classical’ scoring procedures, but the integration of other RS methods (Bottero et al., 2013; Marignani and Blasi, 2012; Sharafi et al., 2012) especially various RS algorithms (e.g. richness complementarity, vegetation and landscape types and multi-criteria analysis) might have an effect on the final PA graphs. Such sensitivity analysis in various conservation areas on a local, regional and global scale could create a general PA graph of each area as a “phenotype” of its biodiversity values. These PA graphs could be useful for comparison of different areas facilitating the conservation planning strategies and priorities. The High-PA introduces areas that should be high priority for conservation. These set of sites combine the highest scores in all the selected criteria and will always be selected regardless of the RS method used. The High-PA index creates a unique ranking of sites (Section 3.2) and can be used as a distinct new scoring procedure for selecting the desired conservation area which combines the specific RS methods used. These areas of high conservation value should become the focus of conservation planning and the top priority on protection initiatives and strategies. The alternative set of sites defined by the Medium-PA of the PAI method creates various conservation solutions. These scenarios of conservation could facilitate the decision makers to introduce to the conservation planning process supplementary socio-economic parameters (Fleishman et al., 2006; Sarkar et al., 2006). In such a complex decision process, the MediumPA offers the ability to introduce external parameters (land ownership, tourism and public

Complimentary Contributor Copy

150

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

distrust of scientists) into a participatory process for avoiding or reducing potential conflicts which could abrogate the total conservation planning project (Gomontean et al., 2008; Margules and Pressey, 2000; Zaccarelli et al., 2008). Finally, the Low-PA, which are areas of low interest for conservation, should become the focus for alternative uses from local communities, especially in areas of undergoing high economic development. Such landsuitability policies at the local and regional level in some cases could succeed to eliminate or alleviate arguments that create land use conflicts between development and conservation (Moilanen and Arponen, 2011). The PAI method and the PA graphs produced that visualize the correlation between the conservation values integration and the potential size of the conservation area could facilitate the understanding of the prioritization process by the stakeholders. This transfer of research knowledge to local communities, if combined with their participation in the selection of the conservation area from a set of alternative conservation scenarios, could establish opportunities for bridging the research-implementation gap on conservation planning and assessment as proposed by Knight et al. (2008).

CONCLUSION The Priority Areas Integration (PAI) method has been developed to facilitate conservation planning by integrating various biodiversity values of a conservation concern area. The PAI method can analyze and visualize the site hierarchization process of multiple reserve selection (RS) methods. On a regional scale, our case study analysis in four SE Aegean islands has created alternatives of conservation rather than defined an optimum solution. Such scenarios could facilitate the planners to design and implement more effective strategies in real conservation planning situations. Finally, a generalized PA graph of the PAI method that analyses and presents a general spatial distribution of the biodiversity values of a conservation concern area is proposed. This unified graph, unique for each area of interest, visualizes the congruencies and variances of these values and based on their integration, produces the alternative conservation scenarios.

ACKNOWLEDGMENTS The authors would like to thank T. Akriotis for his valuable comments and suggestions on the initial manuscript and P. Harris for her linguistic editing which helped present our ideas in English.

REFERENCES Abellan, P., Sanchez-Fernandez, D., Velasco, J., and Millan, A. (2005). Conservation of freshwater biodiversity: A comparison of different area selection methods. Biodiversity and Conservation, 14(14), 3457-3474.

Complimentary Contributor Copy

Priority Areas Integration (PAI) Method

151

Arponen, A., Cabeza, M., Eklund, J., Kujala, H., and Lehtomäki, J. (2010). Costs of integrating economics and conservation planning. Conservation Biology, 24(5), 11981204. Blasi, C., Marignani, M., Copiz, R., Fipaldini, M., Bonacquisti, S., Del Vico, E., et al. (2011). Important Plant Areas in Italy: From data to mapping. Biological Conservation, 144(1), 220-226. Bottero, M., Comino, E., Duriavig, M., Ferretti, V., and Pomarico, S. (2013). The application of a Multicriteria Spatial Decision Support System (MCSDSS) for the assessment of biodiversity conservation in the Province of Varese (Italy). Land Use Policy, 30(1), 730738. Burfield, I. and Van Bommel, F. (2004). Birds in Europe: Population estimates, trends and conservation status. BirdLife Conservation Series No. 12 (p. 374). Cambridge, UK: BirdLife International. Carlstrom, A. (not dated). A survey of the flora and phytogeography of Rhodos, Simi, Tilos and the Marmaris peninsula (SE Greece, SW Turkey): Department of Systematic Botany, University of Lund. Csuti, B., Polasky, S., Williams, P. H., Pressey, R. L., Camm, J. D., Kershaw, M., et al. (1997). A comparison of reserve selection algorithms using data on terrestrial vertebrates in Oregon. Biological Conservation, 80(1), 83-97. Curnutt, J., Lockwood, J., Luh, H. K., Nott, P., and Russell, G. (1994). Hotspots and species diversity. Nature, 367(6461), 326-327. Dimitrakopoulos, P. G., Jones, N., Iosifides, T., Florokapi, I., Lasda, O., Paliouras, F., et al. (2010). Local attitudes on protected areas: Evidence from three Natura 2000 wetland sites in Greece. Journal of Environmental Management, 91(9), 1847-1854. Dimitrakopoulos, P. G., Kalpakas, G. C., Vasios, G. K., and Troumbis, A. Y. (2005). Effectiveness of quantitative methods for selecting priority areas for biodiversity conservation. In: U. Erdem and R. M. Nurlu (Eds.), Proceedings of EURECO 2005 – 10th European Ecological Congress (p. 340). Kusadasi, Turkey. Dimitrakopoulos, P. G., Memtsas, D. and Troumbis, A. Y. (2004). Questioning the effectiveness of the Natura 2000 Special Areas of Conservation strategy: The case of Crete. Global Ecology and Biogeography, 13(3), 199-207. Fleishman, E., Noss, R. F. and Noon, B. R. (2006). Utility and limitations of species richness metrics for conservation planning. Ecological Indicators, 6(3), 543-553. Freitag, S. and Van Jaarsveld, A. S. (1997). Relative occupancy, endemism, taxonomic distinctiveness and vulnerability: Prioritizing regional conservation actions. Biodiversity and Conservation, 6(2), 211-232. Geburek, T., Milasowszky, N., Frank, G., Konrad, H., and Schadauer, K. (2010). The Austrian Forest Biodiversity Index: All in one. Ecological Indicators, 10(3), 753-761. Gomontean, B., Gajaseni, J., Edwards-Jones, G., and Gajaseni, N. (2008). The development of appropriate ecological criteria and indicators for community forest conservation using participatory methods: A case study in northeastern Thailand. Ecological Indicators, 8(5), 614-624. Greuter, W. (1994). Extinctions in Mediterranean areas. Philosophical Transactions - Royal Society of London, B, 344(1307), 41-46. Hernández-Manrique, O. L., Sánchez-Fernández, D., Verdú, J. R., Numa, C., Galante, E., and Lobo, J. M. (2012). Using local autocorrelation analysis to identify conservation areas:

Complimentary Contributor Copy

152

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

An example considering threatened invertebrate species in Spain. Biodiversity and Conservation, 21(8), 2127-2137. Knight, A. T., Cowling, R. M., Rouget, M., Balmford, A., Lombard, A. T., and Campbell, B. M. (2008). Knowing but not doing: Selecting priority conservation areas and the research-implementation gap. Conservation Biology, 22(3), 610-617. Lavergne, S., Molina, J. and Debussche, M. (2006). Fingerprints of environmental change on the rare mediterranean flora: A 115-year study. Global Change Biology, 12(8), 14661478. Lawton, J. H. (1997). The science and non-science of conservation biology. Oikos, 79, 3-5. Lindenmayer, D. B., Steffen, W., Burbidge, A. A., Hughes, L., Kitching, R. L., Musgrave, W., et al. (2010). Conservation strategies in response to rapid climate change: Australia as a case study. Biological Conservation, 143(7), 1587-1593. Lombard, A. T. (1995). The problems with multi-species conservation: do hotspots, ideal reserves and existing reserves coincide? South African Journal of Zoology, 30(3), 145163. Lombard, A. T., Cowling, R. M., Pressey, R. L., and Mustartf, P. J. (1997). Reserve selection in a species-rich and fragmented landscape on the Agulhas Plain, South Africa. Conservation Biology, 11(5), 1101-1116. Margules, C. R., Nicholls, A. O. and Pressey, R. L. (1988). Selecting networks of reserves to maximise biological diversity. Biological Conservation, 43(1), 63-76. Margules, C. R. and Pressey, R. L. (2000). Systematic conservation planning. Nature, 405 (6783), 243-253. Margules, C. R., Pressey, R. L. and Williams, P. H. (2002). Representing biodiversity: Data and procedures for identifying priority areas for conservation. Journal of Biosciences, 27 (4 Suppl. 2), 309-326. Marignani, M. and Blasi, C. (2012). Looking for important plant areas: Selection based on criteria, complementarity, or both? Biodiversity and Conservation, 21(7), 1853-1864. May, R. M., Lawton, J. H. and Stork, N. E. (1995). Assessing extinction rates. In: J. H. Lawton and R. M. May (Eds.), Extinction Rates (pp. 1-24). Oxford: Oxford University Press. Medail, F. and Quezel, P. (1997). Hot-spots analysis for conservation of plant biodiversity in the Mediterranean Basin. Annals of the Missouri Botanical Garden, 84(1), 112-127. Meliadou, A. and Troumbis, A. Y. (1997). Aspects of heterogeneity in the distribution of diversity of the European herpetofauna. Acta Oecologica, 18(4), 393-412. Memtsas, D. P. (2003). Multiobjective programming methods in the reserve selection problem. European Journal of Operational Research, 150(3), 640-652. Memtsas, D. P., Dimitrakopoulos, P. G. and Troumbis, A. Y. (2002). Incorporating multiple ecological criteria in classical zero one selection algorithms. Web Ecology, 3, 48-55. Moffett, A. and Sarkar, S. (2006). Incorporating multiple criteria into the design of conservation area networks: A minireview with recommendations. Diversity and Distributions, 12(2), 125-137. Moilanen, A. (2008). Generalized complementarity and mapping of the concepts of systematic conservation planning. Conservation Biology, 22(6), 1655-1658. Moilanen, A. and Arponen, A. (2011). Setting conservation targets under budgetary constraints. Biological Conservation, 144(1), 650-653.

Complimentary Contributor Copy

Priority Areas Integration (PAI) Method

153

Myers, N. (1990). The biodiversity challenge: expanded hot-spots analysis. Environmentalist, 10(4), 243-256. Oikonomou, V., Dimitrakopoulos, P. G. and Troumbis, A. Y. (2011). Incorporating ecosystem function concept in environmental planning and decision making by means of multi-criteria evaluation: The case-study of Kalloni, Lesbos, Greece. Environmental Management, 47(1), 77-92. Oliver, I. (2002). An expert panel-based approach to the assessment of vegetation condition within the context of biodiversity conservation stage 1: The identification of condition indicators. Ecological Indicators, 2(3), 223-237. Orme, C. D. L., Davies, R. G., Burgess, M., Eigenbrod, F., Pickup, N., Olson, V. A., et al. (2005). Global hotspots of species richness are not congruent with endemism or threat. Nature, 436(7053), 1016-1019. Prendergast, J. R., Quinn, R. M. and Lawton, J. H. (1999). The gaps between theory and practice in selecting nature reserves. Conservation Biology, 13(3), 484-492. Prendergast, J. R., Quinn, R. M., Lawton, J. H., Eversham, B. C., and Gibbons, D. W. (1993). Rare species, the coincidence of diversity hotspots and conservation strategies. Nature, 365(6444), 335-337. Pressey, R. L. and Bottrill, M. C. (2008). Opportunism, threats, and the evolution of systematic conservation planning. Conservation Biology, 22(5), 1340-1345. Pressey, R. L., Humphries, C. J., Margules, C. R., Vane-Wright, R. I., and Williams, P. H. (1993). Beyond opportunism: Key principles for systematic reserve selection. Trends in Ecology and Evolution, 8(4), 124-128. Pressey, R. L. and Nicholls, A. O. (1989). Efficiency in conservation evaluation: Scoring versus iterative approaches. Biological Conservation, 50(1-4), 199-218. Pressey, R. L., Possingham, H. P. and Day, J. R. (1997). Effectiveness of alternative heuristic algorithms for identifying indicative minimum requirements for conservation reserves. Biological Conservation, 80(2), 207-219. Reyers, B., Roux, D. J., Cowling, R. M., Ginsburg, A. E., Nel, J. L., and Farrell, P. O. (2010). Conservation planning as a transdisciplinary process. Conservation Biology, 24(4), 957965. Sarkar, S., Pressey, R. L., Faith, D. P., Margules, C. R., Fuller, T., Stoms, D. M., et al. (2006). Biodiversity conservation planning tools: Present status and challenges for the future. Annual Review of Environment and Resources, 31, 123-159. Sharafi, S. M., Moilanen, A., White, M., and Burgman, M. (2012). Integrating environmental gap analysis with spatial conservation prioritization: A case study from Victoria, Australia. Journal of Environmental Management, 112, 240-251. Similä, M., Kouki, J., Mönkkönen, M., Sippola, A. L., and Huhta, E. (2006). Co-variation and indicators of species diversity: Can richness of forest-dwelling species be predicted in northern boreal forests? Ecological Indicators, 6(4), 686-700. Stork, N. E. (2010). Re-assessing current extinction rates. Biodiversity and Conservation, 19 (2), 357-371. Thuiller, W., Albert, C., Araújo, M. B., Berry, P. M., Cabeza, M., Guisan, A., et al. (2008). Predicting global change impacts on plant species' distributions: Future challenges. Perspectives in Plant Ecology, Evolution and Systematics, 9(3-4), 137-152.

Complimentary Contributor Copy

154

Georgios K. Vasios, Panayiotis G. Dimitrakopoulos and Andreas Y. Troumbis

Troumbis, A. Y. and Dimitrakopoulos, P. G. (1998). Geographic coincidence of diversity threatspots for three taxa and conservation planning in greece. Biological Conservation, 84(1), 1-6. Tucker, G. M. and Heath, M. F. (1994). Birds in Europe: Their conservation status. Birdlife Conservation Series No. 3 (p. 600). Cambridge, UK: BirdLife International. Turland, N. J., Chilton, L. and Press, J. R. (1993). Flora of the Cretan Area: Annotated Checklist and Atlas. London: The Natural History Museum, HMSO. Turnhout, E., Hisschemöller, M. and Eijsackers, H. (2007). Ecological indicators: Between the two fires of science and policy. Ecological Indicators, 7(2), 215-228. Van Jaarsveld, A. S., Freitag, S., Chown, S. L., Muller, C., Koch, S., Hull, H., et al. (1998). Biodiversity assessment and conservation strategies. Science, 279(5359), 2106-2108. Williams, P., Gibbons, D., Margules, C., Rebelo, A., Humphries, C., and Pressey, R. (1996). A comparison of richness hotspots, rarity hotspots, and complementary areas for conserving diversity of British birds. Conservation Biology, 10(1), 155-174. Zaccarelli, N., Riitters, K. H., Petrosillo, I., and Zurlini, G. (2008). Indicating disturbance content and context for preserved areas. Ecological Indicators, 8(6), 841-853.

Reviewed by Dr. Triantaphyllos Akriotis, Assistant Professor at the Department of Environment, University of the Aegean, GR-811 00, Mytilini, Lesvos, Greece (E-mail: takr@ aegean.gr).

Complimentary Contributor Copy

In: National Parks Editor: Johnson B. Smith

ISBN: 978-1-62618-934-8 © 2013 Nova Science Publishers, Inc.

Chapter 5

COMMUNAL GAME FARMING: A SUSTAINABLE LAND USE OPTION? V. Kakembo, P. Mamfengu and G. Kerley Centre for African Conservation Ecology, Nelson Mandela Metropolitan University, Port Elizabeth, South Africa

ABSTRACT Severe land degradation in many communal villages of South Africa has rendered land use forms such as cultivation and livestock farming unviable and unsustainable, making it difficult to earn a land based livelihood. Sustainable communal land management is achievable through the provision of more land use options which are more environmentally friendly. Against the backdrop of land degradation, which is coupled with high poverty levels in the communal villages, game farming has been recommended by ecologists as an ecologically, economically and socially sustainable form of land use. Wildlife conservation in South Africa occurs predominantly at two levels, namely, state owned Parks (National, Provincial and Municipal) and Private Game Reserves. The involvement of local communities has been identified by many researchers as the key to meaningful and effective wildlife conservation. In the Eastern Cape Province, the Tyefu Community Reserve project has been recommended by the Subtropical Thicket Ecosystem Project (STEP), backed by the Department of Environment Affairs and Tourism (DEAT) as a sustainable land use option. Enterprises of this nature have the potential to alleviate poverty whilst conserving biodiversity. However, the following environmental questions arise before such an undertaking is embarked upon: Is the proposed area for the reserve a suitable habitat for wildlife? What kind of wildlife species can be supported, and what is the carrying capacity for the proposed community reserve? How feasible would game farming be in a communal setting against the backdrop of land tenure and land-use constraints? The following aspects are analysed in this case study: the abundance and condition of vegetation, terrain parameters, potential wildlife species and the carrying capacity of the recommended reserve. These should provide the basis for discussion of communal game farming as a land use option in other parts of the country. 

Corresponding author: V. Kakembo. E-mail: [email protected].

Complimentary Contributor Copy

156

V. Kakembo, P. Mamfengu and G. Kerley

1. INTRODUCTION Environmental degradation which manifests itself in soil erosion, a reduction of tree cover in woodlots and the spread of unpalatable species in rangeland is one of the major problems affecting many parts of South Africa. Communal areas, particularly those in the eastern part of the country are characterized by severe soil erosion and vegetation degradation (Boardman et al., 2012). This has rendered cultivation and livestock farming unviable land use options. Many commercial farms, especially in the Eastern Cape Province of South Africa, have switched from agriculture to game farming as an investment and they have realised excellent returns over the last four decades (Smith and Wilson, 2002). In a study undertaken by Langholz and Kerley (2006), it was discovered that some private game farms have seen a 32fold increase in income after conversion. These form prime examples of large areas of land managed for wildlife, with commercial benefits provided by harvesting, live sales and tourism, hence providing an economic incentive for that land use option (Higginbottom and King, 1999). Thus, this form of land use, as opposed to agriculture, is more likely to achieve economical and ecological sustainability with greater benefits for the local communities (Sims-Castley et al., 2004), particularly in the communal areas. Overgrazing by domestic ungulates in the communal areas has extensively transformed wildlife habitat by degrading vegetation, resulting in the loss of phytomass and plant species and the replacement of perennials by annuals (Kerley et al., 1995). However wildlife utilization has been noted as a land use system that, by its nature is less destructive (Brown, 1983). It is often a tourism-based venture aimed at utilizing wildlife resources to benefit the human population (Shackley, 1996). Under appropriate management systems, wildlife conservation can be an economically and ecologically sound alternative to destructive livestock rearing systems in the communal rangelands of South Africa (Brown, 1983). The involvement and support of local communities in wildlife conservation is a prerequisite to effective and long term conservation of wildlife as part of the terrestrial biodiversity (Sibanda and Omwega, 1996). This is particularly true, as only mankind makes decisions which can either enhance the environment or lead to its destruction (Cross, 2001). Wildlife conservation in South Africa occurs predominantly at two levels, namely, state owned Parks (National, Provincial and Municipal) and Private Game Reserves. Community based wildlife conservation is an option that has not taken root. An overview of the two main wildlife conservation levels and the notion of community based reserves is provided below.

2. NATIONAL PARKS In most African countries, wildlife is considered a state resource (Sibanda and Omwega, 1996). In South Africa, officially conserved areas comprise only about 6% of the total land surface (Cross, 2001). Kruger National Park and Kalahari Gemsbok National Park are the premier Reserves of South Africa. There are several private Reserves adjacent to Kruger National Park and they abound throughout the country. Kruger, the largest National Park in the country has more wildlife species than any other game sanctuary (Nolting, 1990). At the time of the Park’s establishment in 1926, the spectacular wildlife assemblage tended to

Complimentary Contributor Copy

Communal Game Farming: A Sustainable Land Use Option?

157

distract attention from other natural resources, and conservation policy became largely focused on game (Bigalke, 2000). The South African National Parks Act of 1926 was promulgated during a period of uncontrolled exploitation of natural resources and Parks were purely concerned with conservation during the first decades (Rupert, 1971). British interests in their African colonies from the eighteenth to nineteenth centuries were mainly concerned with encouraging the mania for game reserves that the ex-hunters and species protectionists desired (Bigalke, 2000). This saw the establishment of the forerunners of several of the major present-day game parks and reserves (Rupert, 1971). The shortcoming of National Parks such as the Kruger has been the exclusion of indigenous people in the establishment of conservation areas (Rupert, 1971). Instead, local communities found themselves having to give up their land to make way for the Park. It was not until the democratic dispensation in 1994 that the Kruger Park started to welcome in its embrace the entire citizenry and other peoples of the world (Cross, 2001). Furthermore, National Parks can no longer consider themselves as remote islands of nature conservation, as Park landscapes extend beyond the Park fence and thus appropriate partnerships with the neighbouring communities have had to be established (Magome and Sondergaard, 2001). Inevitably, the colonial approach of preservation and exclusion gave way to more innovative strategies of resource management and inclusion of local people in the conservation process as key role players (Owen-Smith and Jacobsohn, 1994). Bigalke (2000) points out that conservation agencies now seek to facilitate access of communities living around reserves to a range of natural resources, markets for handcrafts and employment. However, Guyot (2006) argues that the past power relations and stereotypes included in the original national Park idea are a continuing reality.

3. PRIVATE GAME RESERVES In South Africa, the legal rights to utilize wildlife for economic gain have been limited to private and commercial land owners (Brown, 1983; Richardson, 1998). The trend since 1964 is that of a rapidly growing number of privately-owned game and nature reserves. Some of these reserves are of substantial size and very well stocked with game. Most of them have been converted from livestock farming, and cultivation, or a combination of them (SimsCastley et al., 2004; Langholz and Kerley, 2006; Smith and Wilson, 2002). One of the famous examples is Shamwari Game Reserve in the Eastern Cape Province, South Africa, which was created in 1991 through buying and consolidating several overgrazed stock farms. Vegetation clearances to create pastures for sheep, goats and cattle had left the area degraded. As observed by Sims-Castley et al. (2004), the establishment of such reserves has brought about a considerable expanded conservation estate in the Eastern Cape, with apparent benefits in terms of biodiversity conservation. The surrounding communities have benefited directly from such initiatives through employment. It is noteworthy, however, that the present case study is different, as it lies in the communal areas, which did not benefit from such initiatives.

Complimentary Contributor Copy

158

V. Kakembo, P. Mamfengu and G. Kerley

4. COMMUNITY-BASED CONSERVATION The future of wildlife in Africa is threatened, as the largely protectionist approaches to conservation of resources, which have been used for many years, are no longer viable (Sibanda and Omwega, 1996). For example, people are unlikely to support conservation when they are evicted from their homes in order to create protected areas. To avoid this problem, there is a need to create reserves that allow people to harvest resources; to conserve landscapes with long established sustainable resource use patterns and to identify ways of using natural resources without causing environmental degradation (Weddell, 2002). In the present case study, the local communities would gain access to collectively-managed resources of the proposed reserve. Hackel (1999) points out that, with colonialism, conservation policy was based on a coercive form of protectionism that ignored the rights of African people. He emphasizes the need to recognize the involvement of local communities as the key to meaningful and effective wildlife management. The issue of local ownership and control also needs to be taken into consideration. According to Hulme and Murphree (2001), communal game farming is an initiative that aims to make rural people an integral part of the conservation effort. It is another land use option that has the potential of alleviating poverty in rural areas whilst conserving the environment. According to Richardson (1998), community-based wildlife utilization provides an important link between those that bear the costs of wildlife and those who reap the benefits. Kiss (2004) identifies the importance of such projects as the prospect of linking conservation and local livelihoods, preserving biodiversity whilst simultaneously reducing rural poverty. Skyer and Saruchera (2004) observe that access to collectively-managed resources is important for poor rural households. Many governments, however, continue to pursue policies that undermine the livelihoods of those most dependent on these resources by privatisation or entrenching monopoly and state control. In order to ensure maximum benefits for all in community owned wildlife areas, local people have to be involved both in strategic planning and subsequent management, and have to develop a perspective which values wildlife as an economic resource (Shackley, 1996). Conservationists came to realize that local people, who are commonly hostile to wildlife conservation, had to be won over as supporters of their efforts because without their cooperation, wildlife conservation efforts would be doomed (Hackel, 1999; Young, 1998). In southern Africa, several leading community-based natural resource management programmes have been developed. These include Zimbabwe’s Communal Areas Management Programme for Indigenous Resources (CAMPFIRE), Zambia’s Administrative Management Design (ADMADE), Namibia’s National CBNRM Programme, and Mozambique’s Tchuma Tchato project (Skyer and Saruchera, 2004). Sibanda and Omwega, (1996) single out the CAMPFIRE project in Zimbabwe as evidence that communities can be more responsible and more responsive to wildlife conservation when they own wildlife and benefit directly from it. In South Africa, a number of mechanisms have been developed, including contract Park arrangements negotiated between the state and rural communities with land claims in protected areas; joint ventures between communities and the private sector; and providing legal recognition to certain communal property institutions (Skyer and Saruchera, 2004).

Complimentary Contributor Copy

Communal Game Farming: A Sustainable Land Use Option?

159

These approaches have been a great means of involving local communities in their resource management. The present case study falls in the last category of the preceding mechanisms.

5. ENVIRONMENTAL PRE-REQUISITES FOR ESTABLISHING A WILDLIFE CONSERVATION AREA Vegetation Conditions Game farming, as a form of wildlife conservation requires all the requisite habitat ingredients in order to succeed. Each game mammal has its own particular habitat preferences with regard to vegetation type, terrain, and carrying capacity (van Cauter, 2004). Edwards et al. (1997) point out that a common approach is to model animal habitat by linking known habitat use patterns with existing vegetation types, thereby, identifying the extent of important habitat features. Estimates of the abundance and health of wildlife populations can be based on vegetation conditions present including species composition, distribution, and structure (Boshoff et al., 2001). Estimating the potential distribution and abundance of elements of biodiversity is an essential feature of conservation planning (Margules and Pressey, 2000). Different vegetation types offer specific habitats for animals; therefore the distribution patterns of mammals and vegetation often correspond (Turpie and Crowe, 1994). Sustainable management of these ecosystems requires amongst other information, an understanding of the distribution of vegetation types (Schimdt and Skidmore, 2002; Franco-Lopez et al., 2001), especially in response to human activities such as livestock grazing. It is vital to have information on the quality and quantity of vegetation which provides a food source and habitat for wildlife (Goldsmith, 1991; Deal, 1996; Schimdt and Skidmore, 2002). In the present study, the Normalised Difference Vegetation Index (NDVI) is calculated from satellite imagery to determine vegetation condition.

Terrain Parameters The close interaction between vegetation and topography has strong implications for wildlife conservation. Animal movement through habitat terrain is a function of multiple factors including topographic features (Dickson and Beier, 2006). It is necessary to evaluate the habitat of different animals in order to gain an understanding of species-habitat relationships (Nellemann and Fry, 1995). Topography of an area has to provide sufficient local variation to cater for different game animals (Fox et al., 1989). Game animals have different terrain preferences (McCombs II, 1997). For example, some antelopes favour forest and dense vegetation irrespective of the degree of slope whilst other species prefer flat to gently undulating terrain. Flinn et al. (2005) also point out that steeper slopes are less accessible and thus tend to remain forested. Slope and aspect can be singled out as critical topographic controls on vegetation distribution and abundance. Whereas steeper slopes are less accessible and thus tend to be forested, different faunal and floral assemblages can be attributed to differences in

Complimentary Contributor Copy

160

V. Kakembo, P. Mamfengu and G. Kerley

aspect (McCombs II, 1997). Aspect influences the distribution of vegetation in any terrain and in turn, impacts on the distribution of animal species within given terrain (Edwards et al., 1996). Such topographic attributes can be extracted readily from a Digital Elevation Model (DEM). The potential of remote sensing in providing accurate and timely information on essential habitat variables is tremendous. A 20 m DEM is used in the present study to extract slope and aspect terrain parameters.

6. THE TYEFU COMMUNITY RESERVE – A CASE STUDY Against the backdrop of limited land use options, communal game farming has been recommended in the Eastern Cape Province of South Africa by the Subtropical Thicket Ecosystem Planning (STEP) project, supported by the Department of Environmental Affairs and Tourism, as an activity that could promote economic development and poverty alleviation (Sims-Castley et al., 2004), particularly in marginal rural areas with limited agricultural potential (Kiss, 2004), such as the present study area. STEP was funded by the Global Environment Facility through the World Bank to assess the extent of transformation of thicket types and threats facing them (Smith and Wilson, 2002). Experiences from other countries like Botswana and Zimbabwe indicate that revenue generation, local institution and capacity building, and policy reform have been achieved through communal game farming (Hasler, 2004). The Tyefu communal villages in Ngqushwa District, Eastern Cape Province (see Figure 1) are some of the areas where severe erosion, drastic transformation of thicket vegetation and abandonment of eroded arable lands are a widespread phenomenon (Kakembo, 2001; Palmer et al., 2004). The degradation of the natural resource base coupled with limited economic opportunities has resulted in rampant poverty. The area therefore showcases the potential of communal game farming. The steep inter-basin ridges of the Great Fish River catchment, which are adjacent to the Tyefu communal villages, are still covered by dense thicket and semi-succulent thorny scrub. The steep terrain has rendered communally owned land inaccessible for other forms of land use. Parts of the densely vegetated ridges have been identified as the potential location for the proposed Tyefu Community Reserve (TCR). In principle, this land use option would have a lot higher economic returns than other land uses, as demonstrated by privately owned game reserves mentioned earlier. Enterprises of this nature have the potential to alleviate poverty, whilst conserving biodiversity. However, concerns relating to the habitat suitability for wildlife, the kind of wildlife species to be supported, the carrying capacity of the proposed community reserve and the feasibility of game farming arise, before such an undertaking is embarked upon, in view of land use and tenure constraints. Such an undertaking requires an understanding of inter alia, the abundance and condition of vegetation, terrain parameters, and an assessment of potential wildlife species, as well as the carrying capacity of the proposed reserve as a wildlife habitat. The following section assesses the habitat suitability of the proposed TCR as a case study in the light of these parameters. A model of potential wildlife species and carrying capacity of the proposed TCR is also developed. The parameters analysed and model developed in this case study should provide the basis for adopting communal game farming as a land use option in other parts of the country.

Complimentary Contributor Copy

Communal Game Farming: A Sustainable Land Use Option?

161

Figure 1. Study area showing the Tyefu communal villages and location of proposed game reserve.

7. MAPPING VEGETATION AND TERRAIN PARAMETERS FOR THE TCR Mapping of vegetation conditions for the proposed TCR was carried out using the Advanced Spaceborne Thermal Emission and Reflection Radiometer (ASTER) imagery of 15 m resolution taken in March 2009. Bands from the Visible and Near Infrared (VNIR) subsystem, which incorporates three spectral bands covering the wavelength range 0.52-0.86 µm (green, red, and near-infrared) (Lillesand et al., 2004; Poli et al., 2006) were used. They were processed and analysed in Idrisi Andes GIS environment. The Normalised Difference Vegetation Index (NDVI), was calculated using this formula: NDVI = (Infrared – Red) / (Infrared + Red). The NDVI measures the amount of green vegetation based on the principle that actively growing plants absorb radiation in the visible region of te spectrum, while strongly reflecting radiation in the near-infrared region (Ryan, 1997). The NDVI produces values ranging between -1 indicating no vegetation and; +1 indicating complete green vegetation cover (Gibson et al., 2000). Figure 2 is an NDVI image of the proposed TCR and adjacent communal villages.

Complimentary Contributor Copy

162

V. Kakembo, P. Mamfengu and G. Kerley

Figure 2. NDVI image depicting vegetation condition in the proposed reserve and surrounding communal villages. Note the degraded vegetation condition in the Tyefu communal villages to the north of the proposed reserve.

With the exception of a few transformed areas, NDVI values within the proposed reserve range from 0.28 to 1.00, signifying reasonably healthy vegetation cover. A sharp contrast in NDVI values is noticeable between the proposed reserve and communal villages to the north of it. Thicket vegetation in pristine condition was observed during field surveys in most parts of the proposed TCR (see Figure 3). A medium resolution DEM of 20 m was used to calculate terrain parameters such as slope and aspect. It was processed in Idrisi Andes GIS, where attributes of terrain such as slope and aspect were calculated. This would assist in determining habitat suitability and also inform the infrastructural layout of the proposed reserve. A slope surface calculated from the DEM is shown in Figure 4. Whereas slopes range from 10% (1º) to 36% (20º) in the communal villages, the proposed reserve is spanned by a steep ringing ridge with slopes ranging between 44% (24º) and 80% (39º). The slope parameter has played a major role in influencing land use patterns in terms of restricting human activities. It partly explains the existence of healthy thicket vegetation conditions just adjacent to communal villages where vegetation is severely degraded.

Complimentary Contributor Copy

Communal Game Farming: A Sustainable Land Use Option?

163

Figure 3. Part of the proposed reserve showing dense, untransformed thicket vegetation.

The ringing ridge is illustrated in Figure 5, which is a boolean image extracted from the slope surface using ‘reclass’ facility in Idrisi Andes GIS. This could have implications for the infrastructural layout of the proposed reserve, as it could provide excellent vantage points for game viewing. The role of aspect in determining different faunal and floral assemblages, as well as the distribution of animal species was reviewed earlier. Four aspect range surfaces – north-east, south-east, south-west and north-west – were calculated from the DEM in accordance with McCombs II (1997) allocation of aspect ranges for wildlife management. Using the ‘overlay’ module in Idrisi Andes GIS, classified vegetation images were overlaid on the aspect boolean images. The percentage of the different vegetation cover types on the respective aspect ranges was calculated using the ‘database query’ facility in Idrisi Andes GIS. As can be noted from Figure 6, the SW aspect has the highest proportion of thicket vegetation. In the southern hemisphere, animals tend to show a positive selection for northerly and easterly slopes (van Teylingen and Kerley, 1995). McCombs II (1997) attributes this to the increased solar radiation, which can be directly linked to increased floral productivity through greater photosynthetic activity and soil moisture effects, provided there are no other limiting factors. However, owing to the geographical location of the proposed reserve, in close proximity to the east coast, the predominant onshore moist winds make the south east facing slopes moister and hence more thickly vegetated. It is noticeable from Figure 6 that the south facing slopes have more vegetation, particularly dense thicket. Although northerly aspects may be preferred by animals because of their warmth, the southerly facing slopes would create a more desirable feeding area.

Complimentary Contributor Copy

164

V. Kakembo, P. Mamfengu and G. Kerley

Figure 4. Slope surface (%) calculated from the 20 m DEM. Superimposed is the proposed TCR boundary.

8. POTENTIAL SPECIES AND CARRYING CAPACITY ESTIMATION Potential Wildlife Species Potential wildlife species were identified from the distribution estimates based on historic accounts and published literature such as Skead (1987); Mills and Hes (1997); Skinner (1998), Apps (2000) and Skinner and Chimimba (2005) on historical occurrences of wildlife species in the study area. As noted by Boshoff et al. (2001) and Boshoff and Kerley (2001), historical accounts have the following shortcomings: -

Hunters and naturalists recording only species encountered along well traveled routes. Few explorers traveling at night, thus missing nocturnal species. Inaccurate identification of some species. Use of political boundaries for mapping distribution, resulting in generalized ranges of occurrence.

Kerley et al. (2003) point out that ‘potential’ refers to areas where species may be presently absent but could be re-introduced without the need for restoration of habitats. Information required for conserving biological diversity includes amongst other things, the

Complimentary Contributor Copy

Communal Game Farming: A Sustainable Land Use Option?

165

natural distribution ranges and ecological requirements of species (Boshoff and Kerley, 2001). Although these reviews were very useful in determining the general distribution of species, they are frustratingly vague in terms of the exact areas and habitats occupied by the various species (Boshoff and Kerley, 2001). That notwithstanding, species that could be re-introduced into the proposed reserve were identified on the basis of these historic accounts. Table 1 presents a list of these mammals and a summary of their resource requirements such as forage, preferred habitat and topography. Although these species once occurred in the area and could therefore be re-introduced, this habitat has been transformed and thus its potential to support some mammals has been reduced. Furthermore, the small size of the proposed reserve limits the options for some species that require large areas. Hence, some of the species were not included when developing the carrying capacity model for the reserve described below.

Carrying Capacity Model Van Cauter’s (2004) model was used to estimate game distribution and spatial requirements. This was originally developed by Boshoff, Kerley and Cowling (2001) for Baavianskloof Megaconservancy Area (also located in the Eastern Cape Province).

Figure 5. Boolean image of the ringing ridge extracted from the slope surface.

Complimentary Contributor Copy

166

V. Kakembo, P. Mamfengu and G. Kerley

Figure 6. Area covered by different vegetation types on the different aspect ranges.

The model was later modified by Boshoff, Wilson and Skead (2004). This potential herbivore abundance model takes into consideration vegetation types and thus equates them with mammal habitat types. It is based on the productivity estimates of each of the vegetation types and the metabolic requirements of each of the herbivore species (Boshoff et al., 2001). The steps followed in model development were:

a. Allocating Animals into Feeding Guilds Animals were allocated into different feeding guilds based on forage preferences of the herbivores within each vegetation type. The forage guilds include bulk grazers, concentrate grazers, mixed feeders and browsers (Boshoff et al., 2001; van Cauter, 2004). b. Adjustment of Agricultural Stocking Rates The Agricultural Carrying Capacity (ACC) is a concept developed in order to manage domestic livestock production systems (Benjaminsen et al., 2005). It is only suitable for estimating stocking rates of primary grass eaters such as cattle. The ACC, as determined by the Department of Agriculture and Land Affairs is set at 5-9 hectares per Large Stock Unit (LSU) (Nel and Davies, 1999). In the present case study, it was considered not conservative enough for wildlife and was thus adjusted by a certain percentage for each vegetation type. The ACC was divided by the modified carrying capacity which was set relatively high to obtain wildlife carrying capacity. This reduced the estimated carrying capacity to approximately 60% of the agricultural carrying capacity.

Complimentary Contributor Copy

Communal Game Farming: A Sustainable Land Use Option?

167

c. Allocating Forage to Feeding Guilds As observed by Bothma et al. (2004), separating the grazing and browsing components in the diet of wildlife for stocking density calculation ensures that the diversity in the vegetation resources is optimally utilized. Within each vegetation type, the percentage of available forage for each feeding guild was allocated. For each habitat unit, all potential biomass was allocated to the herbivores within each foraging guild, in proportion to the percentage of forage available to each guild (Boshoff et al., 2001). d. Adjusting for Seasonality Habitat requirements for wildlife change during the seasons of the year and vary spatially (van Cauter 2004). When developing a carrying capacity model for the proposed reserve, the seasonality of animal species was determined in order to understand which species are considered ‘Resident’ (found throughout the year), ‘Ephemeral’ (found during certain seasons) or ‘Absent’ (never found) within the different vegetation types in the reserve (Boshoff et al., 2001, 2002; Kerley et al., 2003). Furthermore, some species would be limited to patches within a given vegetation type. Those which had limited patches within a given habitat were also treated as Ephemeral (Boshoff et al., 2001). Densities within the Ephemeral habitats were reduced to 20% of those calculated for Resident habitat (Kerley et al., 2003). e. Calculating Model Outputs The area of each vegetation unit was then divided by hectares per large stock unit to obtain the number of animals that can be supported by the vegetation types within the reserve. The flow diagram (Figure 7) summarizes the steps in developing the carrying capacity model and indicates the values derived.

Adjusted Carrying Capacity A summary of the Adjusted Carrying Capacity (ACC) is presented in Table 2. The adjustment is set at a higher percentage for a vegetation type that is vulnerable to transformation and vice versa (Boshoff et al., 2001; van Cauter, 2004). In the present study, the ACC for the Fish thicket and Fish Valley thicket was adjusted by 60% and 40% respectively. Both thicket types can potentially support 2 bulk grazers, 1 concentrate grazer, 2 mixed feeders and 7 browsers per hectare. The vegetation types found in the reserve are characterized by woody species, hence the higher number of browsers. A summary of the estimation model, listing species that were included in it, which have a suitable habitat and sufficient forage within the proposed reserve, is provided in Table 3. The column ‘present on property’ indicates recommended and omitted species. Five of the species that historically occurred in that area were omitted from the potential community recommended for the proposed reserve because there is insufficient suitable habitat to support viable populations of these species (van Cauter, 2004). For example, black rhinoceros, African buffalo and African elephant were excluded because the proposed reserve, which measures approximately 1800 hectares is too small to

Complimentary Contributor Copy

168

V. Kakembo, P. Mamfengu and G. Kerley

support such mega herbivores. Kerley et al. (2003) points out that large species have large home ranges and low densities, and therefore need large areas with sufficient forage. Table 1. A summary of potential wildlife species based on historical incidence and their habitat requirements Potential Species Blue Duiker

Common Duiker

Springbok

Klipspringer

Steenbok

Kudu

Bushbuck

Red Hartebeest

Eland

Burchell’s Zebra

African Elephant

Buffalo

Forage

Preferred Habitat Confined to forest, thickets or Browser, forest canopy provides its dense coastal bush for forage and staple diet with fallen tree leaves cover Presence of bushes is an essential Mixed feeder, utilising grass and habitat requirement for food and broad leaved forbs for browse shelter Typically a species of arid Predominantly a grazer on short regions, favouring short grass grasses but can browse in dry season. savanna, calcareous pans and dry river beds Mainly a browser feeding on the Restricted to rocky habitat growing shoots, flowers and fruits Exclusively a browser preferring Rare from dense woodlands and forbs, young leaves, flowers and fruits thicket vegetation type Occupies a range of savanna A browser, favouring forbs, creepers, vegetation including dense fruits and pods succulent thickets in Eastern Cape Occurs in forests, closed Primarily a browser but also eats woodlands, thickets and riverine tender grass woodlands

Topography Can occur at altitudes over 1 370m Altitude from sea level to 1 800m above sea level Absent from rocky mountainous areas and rocky hills Can be found on mountain slopes Prefers flatter areas, often absent from rocky mountain slope

Favours rocky hills

Can be found at elevation up to 1 800m

A concentrate grazer

Habitat includes areas grassland

Favour grassy mountain slopes

Mixed feeder, grazes extensively but in winter when green grass is not available may switch to browsing

Highly versatile due to independence from drinking water and ability to utilize a wide variety of food resources. Can be found in arid semi-desert, montane grasslands and forests.

Occurs in altitudes of up to 2 700m. may feed on slopes of up to 350 but prefers those less than 200

A grazer, only taking limited amounts Inhabit open woodland and of browse when the quality and grassland. quantity of available grasses decline. Habitat requirements include dense woodlands, thick Elephants browse and graze, utilizing underbrush and heavy grass cover a wide variety of species for food and shelter as well as sufficient water. Grass and roughage feeders. Mainly Habitat requirements include grazers, but have been observed abundant grass, shade and water browsing on rare occasions

Resident on plateaus and plains

Flat to undulating terrain

Prefer flatter or flood plains

Skead, 1987; Skinner and Chimimba, 2005; Mills and Hes, 1997; Skinner, 1998; Apps, 2000.

Other species such as Burchell’s zebra were also excluded from the recommended list because it is a concentrate grazer, yet vegetation in the area is predominantly thicket

Complimentary Contributor Copy

Communal Game Farming: A Sustainable Land Use Option?

169

vegetation (Skinner and Chimimba 2005). Steenbok are rare from dense woodlands and thicket vegetation type (Mills and Hes, 1997), hence their exclusion from the model as well. Three of the recommended game species namely, Common duiker, Kudu and Bushbuck are 1 amongst the top ten popular species hunted by foreign hunters.

Area Covered by each vegetation type

Fish Valley Thicket = 958. 3 ha*

Bulk Grazers = 25% Concentrate Grazers = 5% Mixed Feeders = 15% Browsers = 55%

Fish Thicket = 870.9 ha*

Forage allocation

Agric CC = 5 ha per LSU Adjusted by 55% Modified CC = 7 ha per LSU

*958.3 / 3.01 (ha per LSU) = 318.2 (number of LSUs per habitat type)

Bulk Grazers = 20% Concentrate Grazers = 5% Mixed Feeders = 5% Browsers = 70%

Agric CC = 9 ha per LSU Adjusted by 60% Modified CC = 13 ha per LSU

* 870.9 /15.01 (ha per LSU) = 58.0 (number of LSUs per habitat type)

Figure 7 A flow chart summarizing steps in developing the carrying capacity model with data. Figure 7. A flow chart summarizing steps in developing the carrying capacity model with data. (After (After2004). Van Cauter 2004). Van Cauter

Model Outputs The model outputs are expressed as individuals per hectare, as well as total number of individuals per habitat. The individual per hectare column in Table 3 shows how many animals can be supported by the available land covered by a particular vegetation type). For example, the Fish Thicket covers about 870.9 hectares in the proposed reserve and about 15 hectares of this vegetation type per LSU are required (see Table 2). The model output then shows how many of the individual species Fish Thicket can support. The proposed reserve is capable of supporting a total of 1950 individual animals, of which most are small antelopes. Only a total of seven different types of wildlife species can be supported. Although the model has its limitations, it provides a useful guide to habitat preferences of different herbivores that can potentially be supported by the reserve.

Complimentary Contributor Copy

170

V. Kakembo, P. Mamfengu and G. Kerley Table 2. A Summary of adjusted carrying capacity and feeding guilds within the different vegetation types

Fish Thicket. Agricultural carrying capacity Carrying capacity adjustment Modified carrying capacity: Area of unit: Number of LSU's:

0.111 60 0.0666 870.9 ha 58.0

Feeding Guilds: Bulk grazers Concentrate grazers Mixed feeders Browsers

Potential 2 1 2 7

Fish Valley Thicket Agricultural carrying capacity Carrying capacity adjustment Modified carrying capacity: Area of unit: Number of LSU's:

0.83 40 0.332 958.3 318.17 ha

Feeding Guilds: Bulk grazers Concentrate grazers Mixed feeders Browsers

Potential 2 1 2 7

LSU's per hectare % of Agricultural carrying capacity 15.01 (hectares per LSU)

LSU's per hectare % of Agricultural carrying capacity 3.01 (hectares per LSU)

9. FEASIBILITY OF GAME FARMING IN A COMMUNAL SETTING The feasibility of game farming in a communal setting, coupled with land degradation constraints is another question this case study seeks to answer. The ecologic, economic and social sustainability implications of the proposed TCR are reviewed in this section.

Ecologic Sustainability Game farming has been touted for its potential to allow sustainable development of ecosystems and provide incentives for the conservation of wildlife (Isaacs, 2000; Castley et al., 2001). The proposed Tyefu Communal Reserve, particularly for hunting purposes has the potential of generating incentives for biodiversity conservation. One disadvantage, however, of such an initiative is that it is very dependent on the preferences of the consumers who are often ignorant of the benefits of the natural resources to the community (Isaacs, 2000) and hunters are willing to pay large sums of money to hunt animals (Coltman et al., 2003).

Complimentary Contributor Copy

171

Communal Game Farming: A Sustainable Land Use Option? Table 3. A summary of the estimation model

Herbivore

Species

LSU Foraging equivalents Guild

African elephant* Burchell's zebra*

Loxodonta africana

2.78

Equus burchelli 0.66

mixed feeder

Seasonality Present on Fish Valley property Fish Thicket Thicket no

R

E

bulk grazer no

A

E

Sylvicapra 0.09 browser yes R R grimmia Oreotragus Klipspringer 0.07 browser yes A E oreotragus Raphicerus Steenbok* 0.06 browser no E A campestrus African buffalo* Syncerus caffer 1.07 bulk grazer no E R Tragelaphus Kudu 0.54 browser yes R R strepiceros Tragelaphus mixed Eland 1.08 yes E E oryx feeder Tragelaphus Bushbuck 0.13 browser yes R R scriptus Black Diceros 1.65 browser no E R * rhinoceros bicornis Alcelaphus concentrate Red hartebeest 0.37 yes A R buselaphus grazer cephalophus Blue duiker 0.09 browser yes R A monticola * Species that were omitted from the model. Seasonality/Patchiness: R = Resident (Mammal found in habitat type throughout the year). E = Ephemeral (Mammal found on certain seasons). A = Absent (Mammal never occurs in that habitat). Common duiker

However, if planned well and managed properly, the proposed reserve can enhance sustainable resource use and wildlife management. The benefits of protecting biodiversity include habitat protection, soil formation and nutrient recycling (Isaacs, 2000). It is suggested that local community members should also be allowed in the reserve occasionally to harvest natural resources, thereby promoting sustainable use of resources.

Economic Sustainability Tyefu local communities face challenges, particularly the degraded and unproductive agricultural land, strong reliance on social welfare grants, unemployment and thus widespread poverty. The proposed reserve presents an opportunity to generate income for them. According to Emerton and Mfunda (1999), game farming is a potential mechanism for spreading wealth from affluent to poorer rural communities. Through income generated from live game sales, venison and trophy hunting, as well as other activities such as the provision

Complimentary Contributor Copy

172

V. Kakembo, P. Mamfengu and G. Kerley

of game viewing opportunities, the community can benefit immensely from wildlife. In addition, job opportunities (casual and permanent), support for tourism related small enterprise development, and upgrading of infrastructure by Local Government all accrue from this enterprise. Other benefits as envisaged by Jones (1999), comprise increased skills and local institutional capacity, better interest and pride in culture, assets and identity of an area, and improved local management of natural resources. Notwithstanding the benefit from game farming as indicated by Emerton and Mfunda and Jones in the late 1990s, it is noteworthy that in North America, it is very difficult to balance conservation and income generation (Freese and Trauser, 2000). For example, many wildlife operations appear to be directed by short-term commercial benefits rather than by long term conservation goals, as exemplified by the number of species introduced in some game reserves for purposes of attracting hunters (Castley et al., 2005; van Niekerk, 2006). Therefore, there is a need to sensitise the Tyefu local communities to these trade-offs between short term benefits and long term conservation goals. Such sensitization would be accomplished by organizing dedicated community workshops during the pre and post- reserve implementation phases.

Social Sustainability The communal setting of the Tyefu communal villages is characterized by limited tenure security and lack of clarity in terms of ownership, which often discourages conservation and improvement of natural resources (Metcalfe, 1999). If the local community does not hold a title to the reserve, they might not develop a sense of ownership of such an initiative, thus making it unsustainable. As pointed out earlier, the proposed area for the reserve is small, relative to the adjacent privately owned game reserves and thus more land is required for its expansion. However, the problem with the communal land is the long and difficult process of accessing it from the Department of Land Affairs (Ngqushwa Reviewed IDP, 2006/7). Therefore, tenure challenges are not only about ownership, but also access and rights to a range of renewable and nonrenewable resources (Metcalfe, 1999). One advantage about the proposed reserve is its nondisplacement of existing land use forms; it provides another land use option with far greater benefits. By implication, there will be no land use conflicts. Notwithstanding the access constraints pointed out above, it is recommended that the proposed reserve area be stretched to link up with the nearby Great Fish River Reserve Complex. Only seven wildlife species have a suitable habitat within the proposed reserve. Since hunting has been pointed out above as a feasible activity in the reserve, there is a need for a greater variety of species. Expanding the reserve would ensure more forage and habitat preference, thus greater capacity to support a larger variety of species.

CONCLUSION This chapter has reviewed the wildlife conservation scenario in South Africa. The exclusion of indigenous people in the establishment and management of conservation areas,

Complimentary Contributor Copy

Communal Game Farming: A Sustainable Land Use Option?

173

particularly National Parks has been highlighted. Local communities had to give up their land to make way for Park establishment and expansion. This colonial approach of preservation and exclusion inevitably gave way to the inclusion of local people in the conservation process since the dawn of the 1994 democratic dispensation. The conversion from agriculture to game farming has been the predominant trend in the creation of private game reserves. Mechanisms to provide legal recognition to certain communal property institutions; partnerships between the state and rural communities with land claims in protected areas, as well as communities and the private sector have been developed. However, community based wildlife conservation is an option that has not taken root in South Africa. The habitat suitability of the proposed TCR in the light of vegetation and terrain parameters, potential wildlife species and the carrying capacity of the proposed reserve as a wildlife habitat have been assessed. Vegetation mapping revealed a distinct distribution pattern; the dense thicket that covers most of the proposed reserve has the potential to provide forage and suitable habitat for different wildlife species. The topographic variables of slope and aspect have strong implications for the potential distribution of animals across the proposed reserve. Different game species favour varying degrees of slope and the ringing ridge identified could determine the infrastructural layout of the proposed reserve and provide excellent vantage points for game viewing. The importance of aspect as an important determinant for different faunal and floral assemblages has been elucidated. The moist south facing slopes of the proposed reserve tend to have more vegetation than the warm dry north facing slopes. Despite the tendency for animals to show a preference for northerly and easterly slopes, south facing slopes of the case study area would be preferable for most game animals. This understanding would give an idea as to the relevant hunting and game viewing sites. The wildlife Carrying Capacity Model developed provides a link between forage (vegetation) and habitat (topography). The potential wildlife species for the proposed reserve were identified on the basis of forage availability, whilst the abundance of species was determined on the basis of size of the suitable habitat in the proposed reserve. The total number of individual animals and different wildlife species that can be supported by the reserve was determined on this basis. The ecological, economic and social sustainability of communal game farming lies in balancing conservation and income generation. Short-term commercial benefits should not be pursued at the expense of long-term conservation imperatives. There is a need to develop a sense of security of tenure among the communal local communities. Tyefu local communities should be allowed to hold title to the proposed reserve in their own names. With a sense of property ownership, this will ensure total commitment to the long-term conservation goals. Given its limited areal extent, it is recommended that the proposed reserve be stretched to link up with the nearby Great Fish River Reserve Complex, as this will ensure more forage and habitat preference, as well as greater capacity to support a larger variety of species. This process should, however, be preceded by exhaustive community consultations.

Complimentary Contributor Copy

174

V. Kakembo, P. Mamfengu and G. Kerley

REFERENCES Apps, P. (ed.). (2000) Smither’s Mammals of Southern Africa: A Field Guide. Struik Publishers: Cape Town. Benjaminsen, P. A., Rhode, R., Sjaastad, E., and Lebert, T. (2005) The Politics of Land and Livestock. Available on-line: http://www2.siu.no/vev.nsf/o/siu. Bigalke, R. C. (2000) Functional Relationships between Protected and Agricultural Areas in South Africa and Namibia. In: Prins, H. H. T., Grootenhuis, J. G. and Dolan, T. T. (2000) Wildlife Conservation by Sustainable Use. Kluwer Academic Publishers: London. Boardman, J., Hoffman, T., Holmes, J., and Wiggs, F. S. (2012) Soil Erosion and Land Degradation, In: Holmes, P. and Meadows, M. (Eds). Southern African Geomorphology, Sun Press, Bloemfontein, 305-328. Boshoff, A. F., Kerley, G. I. H. and Cowling, R. M. (2001) A Pragmatic Approach to Estimating the Distributions and Spatial Requirements of the Medium-to Large-sized Mammals in the Cape Floristic Region, South Africa. Diversity and Distributions, 7: 29-43. Boshoff, A. F., Kerley, G. I. H. and Cowling, R. M. (2001) A Pragmatic Approach to Estimating the Distributions and Spatial Requirements of the Medium-to Large-sized Mammals in the Cape Floristic Region, South Africa. Diversity and Distributions, 7: 29-43. Boshoff, A. F. and Kerley, G. I. H. (2001) Potential Distributions of the Medium-to-Largesized Mammals in the Cape Floristic Region, Based on Historical Accounts and Habitat Requirements. African Zoology 36(2): 245-273. Boshoff, A. F., Kerley, G. I. H. and Cowling, R. M. (2002) Estimated Spatial Requirements of the Medium-to Large-sized Mammals, according to Broad Habitat Units, in the Cape Floristic Region, South Africa. African Journal of Range and Forage Science, 19:29-44. Boshoff, A. F., Wilson, S. L. and Skead, C. J. (2004) Medium-to Large-sized Mammalian Herbivores in Provincial Nature Reserves in the Eastern Cape Province: Broad Habitat Types, Potential Species, Inappropriate Species and Selected Management Recommendations. TERU. University of Port Elizabeth, Report C106. Bothma, J., Du, P., Van Rooyen, N., and Van Rooyen, M. W. (2004) Using Diet and Plant Resources to set Wildlife stocking densities in African Savannas. Wildlife Society Bulletin, 32(3): 840-851. Brown, E. K. (1983) Wildlife: A paying proposition? In Which Way Botswana Wildlife?: Proceedings of a Symposium Kalahari Conservation Society 1983. P51. Castley, J. G., Boshoff, A. F. and Kerley, G. I. H. (2001) Compromising South Africa’s natural biodiversity-inappropriate herbivore introductions. South African Journal of Science, 97. 344-348. Coltman, D. W., O’Donogghue, P., Jorgenson, J. T., Hogg, J. T., Strobeck, C., and FestaBianchet, M. (2003) Undesirable Evolutionary Consequences of Trophy Hunting. Nature, 426: 655-658. Cross, R. (ed). (2001) The Wildlife and Environment Society of South Africa: Celebrating 75 years of people caring for the Earth. Wessa: South Africa. Deal, K. H. (1998) Wildlife and Natural Resource Management. New York: Delmar Publishers.

Complimentary Contributor Copy

Communal Game Farming: A Sustainable Land Use Option?

175

Dickson, B. G. and Beier, P. (2006) Quantifying the Influence of Topographic Position on Cougar (Puma Concolor) Movement in Southern Carlifornia, US. Journal of Zoology. Emerton, L. and Mfunda, I. (eds.) (1999) Making Wildlife Economically Viable for Communities Living around the Western Serengeti, Tanzania. Biodiversity Economics for Eastern Africa, IUCN: Kenya. Flinn, K. M., Vellend, M. and Marks, P. L. (2005) Environmental Causes and Consequences of Forest Clearance and Agricultural Abandonment in Central New York, US. Journal of Biogeography, 32: 439-452. Fox, J. L., Smith, C. A. and Schoen, J. W. (1989) Relationship between Mountain Goats and their Habitat in Southern Alaska. Gen. Tech. Rep. PNW-GTR-246: US Department of Agriculture, Forest Service, Pacific Northwest Research Station. 25p. Franco-Lopez, H., Ek, A. R. and Bauer, M. E. (2001) Estimating and Mapping Forest Stand Density, Volume and Cover Type using th K-nearest Neighbors Method. Remote Sensing of Environment. 77: 251-274. Freese, C. H. and Trauser, D. L. (2000) Wildlife Markets and Biodiversity Conservation in North America. Wildlife Society Bulletin, 23(1): 114-119. Gibson, P. J. and Power, C. H. (2000) Introductory Remote Sensing: Digital Image Processing and Applications. Taylor and Francis group: New York Goldsmith, F. B. (1991) Monitoring for Conservation and Ecology. Chapman and Hall, London. Guyot, S. 2006. Review of Ramutsindela, Maano, Parks and People in Postcolonial Societies: Experiences in Southern Africa. HSAfrica, H-Net Reviews. Hackel, J. D. (1999) Community Conservation and the Future of Africa’s. Wildlife Conservation Biology, 13(4): 726-734. Hasler, R. (2004) The Institutional Paradox of Community Based Wildlife Management: Project for Land and Agrarian Studies. University of the Western Cape. Higginbottom, K. and King, N. (1999) The Live Trade in Free-ranging Wildlife within South Africa, and Implications for Australia. A Report for the Rural Industries Research and Development Co-operation. RIRDC Publication number: 06/046. Hulme, D. and Murphree, M. (2001) African Wildlife and Livelihoods: The Promise and Performance of Community Conservation. James Currey Ltd: Great Britain. Isaacs, J. C. (2000) The Limited Potential of Ecotourism to Contribute to Wildlife Conservation. Wildlife Society Bulletin, 28(1): 61-69. Jones, B. T. B. (1999) Rights, Revenue and Resources: The Problems and Potential of Conservancies as Community Wildlife Management Institutions in Namibia. Evaluating Eden Series. Discusion Paper No. 2. Kakembo, V. (2001) Trends in Vegetation Degradation in relation to Land Tenure, Rainfall, and Population changes in Peddie District, Eastern Cape: South Africa. Environmental Management, 28(1): 39-46. Kerley, G. I. H., Knight, M. H. and De Kock, M. (1995) Desertification of Subtropical Thicket in the Eastern Cape, South Africa: Are There Alternatives? Environmental Monitoring and Assessment, 37(1-3): 211-230. Kerley, G. I. H., Pressey, R. L., Cowling, R. M., Boshoff, A. F., and Sims-Castley. (2003) Options for the Conservation of Large and Medium-sized Mammals in the Cape Floristic Region Hotspot, South Africa. Biological Conservation, 112: 169-190.

Complimentary Contributor Copy

176

V. Kakembo, P. Mamfengu and G. Kerley

Kiss, A. (2004) Is community-based ecotourism a good use of biodiversity conservation funds? Trends in Ecology and Evolution, 19(5): 232-237. Langholz, J. A. and Kerley, G. I. H. (2006) Combining Conservation and Development on Private Lands: An Assessment of Ecotourism-based Private Game Reserves in the Eastern Cape. Centre for African Conservation Report No. 56. Lillesand, T. M., Kiefer, R. W. and Chipman, J. W. (2004) Remote Sensing and Image Interpretation. John Wiley and Sons Incl: US. Margules, C. R. and Pressey, R. L. (2000) Systematic Conservation Planning. Nature, 405: 243-253. McCombs II, J. W. (1997) Geographic Information System Topographic Factor Maps for Wildlife Management. MSc Theses, Polytechnic Institute and State University. Virginia. Metcalfe, S. (1999) Natural Resource Tenure in the Context of Sustainable Use. IUCN, Gland, Switzerland and Cambridge, UK. pp. 31-42. Mills, G. and Hes, L. (1997) The Complete Book of Southern African Mammals. Struik Publishers: Cape Town. Nel, E. L. and Davies, J., (1999) Farming Against the Odds: An Examination of the Challenges Facing Farming and Rural Development in the Eastern Cape Province of South Africa. Applied Geography, 19: 253-274. Nellemann, C. and Fry, G. (1995) Quantitative Analysis of Terrain Ruggedness in Reindeer Winter Grounds. Arctic, 48(2): 172-176. Ngqushwa Municipality Reviewed Integrated Development Plan (IDP), (2006/07). Eastern Cape, South Africa. Nolting, M. W., (1990) Africa’s Top Wildlife Countries. Global Travel Publishers, inc: US. Owen-Smith, G. and Jacobsohn, M. (1994) Namibia: Integrated Rural Development. In: Vision of Wildlife, Ecotourism and the Environment in Southern Africa: Endangered Wildlife Trust. Vision annual 1994. P138. Palmer, A. R., Kakembo, V., Lloyd, J. W., Ainslie, A. (2004) Degradation patterns in the succulent thicket. Proceedings of the First STEP Forum, Zuurberg, South Africa. Poli, D., Remondino, F. and Dolci, C. (2006) Use of Satellite Imagery for DEM extraction, Landscape Modelling and GIS Applications. Switzerland: Institute of Geodesy and Photogrammetry. Richardson, J. A. (1998) Wildlife utilization and biodiversity conservation in Namibia: conflicting or complementary objectives? Biodiversity and Conservation, 7: 549-559. Rupert, A. (1971) Management in Nature Conservation. Published by the South African Wildlife Foundation Magome and Sondergaard, 2001). Ryan, L. (1997) Creating a ‘Normalized Difference Vegetation Index’ NDVI Image Using MultiSpec. University of New Hempshire, Durham. Schmidt, K. S. and Skidmore, A. K. (2002) Spectral Discrimination of Vegetation Types in a Coastal Wetland. Remote Sensing of Environment, 85: 92-108. Shackley, M. (1996) Wildlife Tourism. London: International Thompson Business Press. Sibanda, B. M. C. and Omwega, A. K. (1996) Some reflections on conservation, sustainable development and equitable sharing of benefits from wildlife in Africa: the case of Kenya and Zimbabwe. Wildlife Research, 26(4): 175-181. Sims-Castley, R., Kerley, G. I. H. and Geach, B. (2004) A Questionnaire Based Assessment of the socio-economic significance of ecotourism-based private game reserves in the Eastern Cape. Report No. 51. Terrestrial Ecology Research Unit, Port Elizabeth.

Complimentary Contributor Copy

Communal Game Farming: A Sustainable Land Use Option?

177

Skead, C. J. (1987) Historical Mammal Incidence in the Cape Province, Volume 2: The Eastern Half of the Cape Province, Including the Ciskie, Transkie and East Griqualand. Chief Directorate Nature and Environmental Conservation, Cape Town. Skinner, J. (1998) Struik Pocket Guide: Mammals of Southern Africa: A Field Guide. Struik Publishers, Cape Town. Skinner, J. D. and Chimimba, C. T. (2005) The Mammals of the Southern African Subregion. Cambridge University Press: Cape Town. Skyer, P. and Saruchera, M. (2004) Community Conservancies in Namibia: An Effective Institutional Model for Commons Management? Cape Town: Programme for Land and Agrarian Studies, University of the Western Cape. PLAAS Policy Brief 14. Smith, N. and Wilson, S. (2002) Changing Land-use Trends in the Thicket Biome: Pastoralism to Game Farming. TERU Report No. 38. University of Port Elizabeth, South Africa. STEP (2004) Fish River Biodiversity Initiative. Available on-line: http://bgis.sanbi.org/STEP/ index.asp. Turpie, J. K. and Crowe, T. M. (1994) Patterns of Distribution, Diversity and Endemism of Larger African Mammals. South African Journal of Zoology, 29(1): 19-32. Van Cauter, A. (2004) Developing and Testing Models of Mammal Community Structure in the Eastern Karoo. MSc Thesis. University of Port Elizabeth, Port Elizabeth. Van Cauter, A., Kerley, G. I. H. and Cowling, R. M. (2005) The Consequences of Inaccuracies in Remote-Sensed vegetation Boundaries for Modelled Mammal Population Estimates. South African Journal of Wildlife Research 35(2): 155-161. Van NIekerk, Du, P. P. (2006) Characteristics of the Game Industry in the Eastern Cape: A Brief Overview. In: Wilson, S. L. (ed). Proceedings of the 2004 Thicket Forum. Center for African Conservation Ecology, Report No. 4, Nelson Mandela Metropolitan University, South Africa. Weddell, B. J. (2002). Conserving Living Resources: In the context of a changing world. United KIngdom: Cambridge University Press. Young, A. (1998) Land Resources: Now and for the future. Cambridge university press: Cambridge.

Complimentary Contributor Copy

Complimentary Contributor Copy

In: National Parks Editor: Johnson B. Smith

ISBN: 978-1-62618-934-8 © 2013 Nova Science Publishers, Inc.

Chapter 6

LAND AS SUSTENANCE AND SANCTUARY: SETTLEMENT HISTORY AND RESOURCE USE IN AND AROUND UTAH’S GRAND STAIRCASE-ESCALANTE NATIONAL MONUMENT Robert J. Lilieholm1, and Marietta Eaton2,† 1

E. L. Giddings Associate Professor of Forest Policy, School of Forest Resources, The University of Maine, Orono, ME, US 2 Canyons of the Ancients National Monument Manager and Anasazi Heritage Center Director, Bureau of Land Management, National Landscape Conservation System, Dolores, CO, US

ABSTRACT President Bill Clinton designated the 1.9-million-acre Grand Staircase-Escalante National Monument (GSENM or Monument) in September of 1996. While the Monument’s creation was largely intended to protect the region’s scientific and historic values, the growing economic importance of recreation and tourism was an important though less-noted factor. In this respect, the establishment of GSENM represents the culmination of protection efforts and counter-efforts in the region that date back to the 1930s – when parts of today’s Monument were first considered for federal protection within the USDI National Park Service. The prolonged struggle to preserve the area’s remoteness has sparked a string of controversies – a feature of the land and culture that will likely endure. This chapter describes the social, economic and cultural history of the people that settled in and around GSENM in an effort to provide context for past, current, and future debates surrounding the Monument’s creation and management. The lessons



Corresponding author: Robert J. Lilieholm. E.L. Giddings Associate Professor of Forest Policy, School of Forest Resources, The University of Maine, Orono, ME 04469-5755. Phone: (207) 581-2896; Fax: (207) 581-2875; E-mail: [email protected]. † Marietta Eaton. Canyons of the Ancients National Monument Manager and Anasazi Heritage Center Director, Bureau of Land Management, National Landscape Conservation System, 27501 Hwy 184, Dolores, CO 81323.

Complimentary Contributor Copy

180

Robert J. Lilieholm and Marietta Eaton learned here provide useful insights for other regions of the globe seeking to balance resource protection and changing social demands.

Keywords: Agriculture, economics, ecotourism, environmental policy, grazing, mining, Mormons, parks and protected areas, ranching, recreation, resource-dependent communities, rural development

“The Grand Staircase-Escalante National Monument’s vast and austere landscape embraces a spectacular array of scientific and historic resources. This high, rugged, and remote region, where bold plateaus and multi-hued cliffs run for distances that defy human perspective, was the last place in the continental United States to be mapped. Even today, this unspoiled natural area remains a frontier, a quality that greatly enhances the monument’s value for scientific study. The monument has a long and dignified human history: it is a place where one can see how nature shapes human endeavors in the American West, where distance and aridity have been pitted against our dreams and courage.” President Bill Clinton, September 18, 1996

INTRODUCTION The Grand Staircase-Escalante National Monument (GSENM or Monument) encompasses nearly two million acres of rugged, isolated desert in south-central Utah (Figure 1). With an elevational range of nearly 6,000 feet, the GSENM includes a diversity of environments, including high plateaus and mesas, sweeping deserts, incised canyons, and rolling sandstone slickrock (Figure 2). The Monument provides a physiographic delineation of three major zones. The western zone – the “Grand Staircase” – was named by geologist Clarence Dutton in 1880 for its series of brightly hued cliffs that form an elevational and chronological gradient from Bryce Canyon National Park (elevation 9,115 feet), located north of the Monument, to the depths of the Grand Canyon (elevation ~2,200 feet) located to the Monument’s south (Malakoff 1999). The Kaiparowits Plateau forms the central zone of the GSENM – an upland area of cretaceous-age benches, cliffs, and geological formations. The last zone, the Escalante Canyons, contains deeply incised river and tributary canyons, along with upland desert. In this land of climatic extremes, where intermittent streams, springs, and hidden water pockets determine survival, humans have carved an existence from the land for centuries. Even now, the Monument’s vastness and aridity commonly mask the centuries-old human imprint on the land, and conditions of geography, drought, and flash-floods challenge survival. Today, scattered within this vast region lie a handful of small communities known for their devout Mormon faith and near-legendary history of conflict with federal authorities over the use and management of public lands. Indeed, aside from the region’s spectacular and austere beauty, southern Utah is perhaps best known for its vocal opposition to the federal land management agencies that oversee 66% of the State.

Complimentary Contributor Copy

Land As Sustenance and Sanctuary

181

Figure 1. Utah’s Grand Staircase-Escalante National Monument.

Nowhere has this battle been more intense than in Kane and Garfield Counties – home to GSENM and a region where public lands comprise more than 90% of the land base. The connections between land and people, religion and community, and local versus national interests in public land are inextricably woven throughout the post-settlement history of the region. These relationships are of long-standing interest to academics and land managers alike, and often set the bounds and determine the success of public land policies (Foltz 2000). This chapter explores the region’s exploration and settlement in order to provide historical context to the contentious land use issues of today. We then use this background to identify core features of GSENM communities that influence community-resource perspectives, and comment on the opportunities and challenges these views pose for managing public lands in this region of growing national and international significance. Although the area in and around GSENM has been home and passageway for millennia to Native Americans (Figure 3), this paper focuses mostly on the European-American (EuroAmerican) influence on and perceptions of the landscape.

Complimentary Contributor Copy

182

Robert J. Lilieholm and Marietta Eaton

Photo credit: BLM. Figure 2. Lower Calf Creek Falls, GSENM.

EXPLORATION, SETTLEMENT AND RESOURCE USE Spanish Explorers and the Arrival of Mormon Settlers The Dominguez/Escalante Expedition of 1776 represents the first Euro-American foray into the GSENM region. Although the expedition, seeking a route to the California missions, crossed the Colorado Plateau several miles south of the present-day Monument, the Jesuits played an important role in opening the region to further exploration. By 1827, Antonio Armijo was among the first of a number of traders seeking a more direct route between Santa Fe and California – a route that would later be called the Old Spanish Trail. His expedition actually traversed the southern boundary of the Monument, roughly following the current alignment of Highway 89 between Kanab, Utah, and Page, Arizona (Figure 1). Life for the region’s Native Americans – Southern Paiute, Ute, Navajo and Hopi – was little affected by Spanish incursions, and many tribes were able to continue traditional lifestyles. The Southern Paiute and a few Ute and Navajo used portions of the Monument, practicing traditions that had evolved over centuries. The Navajo conducted forays into Paiute territory, stealing slaves for the lucrative trade in California and Santa Fe.

Complimentary Contributor Copy

Land As Sustenance and Sanctuary

183

Photo credit: BLM. Figure 3. Anasazi ruin, Flood Canyon, GSENM.

In 1847, the first members of the Church of Jesus Christ of Latter Day Saints (the Church, Saints, LDS, or Mormons) arrived in the Great Salt Lake Valley – 200 miles north of the Monument. The severity and isolation of Deseret – the name given to the region by LDS church president Brigham Young – was seen as a place to escape not only persecution, but also as a refuge from mainstream American culture. The region allowed seclusion and protection, a refuge from unwanted social change, a sanctuary from non-LDS “Gentiles” and a world with values at-odd with Mormon culture (Jackson 1978). Indeed, Mormon doctrine blurred the separation of church and state, and in budding communities the general tenet put the community welfare above that of the individual. The settling of Utah was therefore considered to be a divine task, the destined building of Deseret along the Mormon Corridor that was envisioned to one day reach from Salt Lake City west to the shores of southern California. By 1851, the Church began a systematic program of settlement in southern Utah, defining locations and specific “missions” for new communities. Parowan, Paragonah, and Cedar City were settled first. Cedar City, with its rich mineral deposits, was established as the “Iron Mission.” Soon the area around St. George was settled as the “Cotton Mission” focused on farming. Exports to the rest of Deseret revolved around minerals, cotton, livestock, and dairy products. As unclaimed land and water in the Cedar City and St. George areas became scarce, settlers began to move outward in search of rangelands and well-watered areas for farming and ranching. A Mormon ranch was established at Pipe Spring, Arizona, in 1866, followed by

Complimentary Contributor Copy

184

Robert J. Lilieholm and Marietta Eaton

a series of mission settlements surrounding the Monument between 1863 and 1876. The strongest expression of the Mormon’s colonizing skills emerged in Kanab along the wellwatered reaches of Kanab Creek. Additional settlements were established in Long Valley along the East Fork of the Virgin River at Berryville (modern Glendale), Orderville, Winsor (modern Mt. Carmel), and Hatch. Other settlements included Upper Kanab (also known as Graham and ultimately Alton), which was settled, abandoned, and then re-occupied. The original settlement of Pahreah, located within GSENM, was established by Peter Shirts, who constructed a small fortress south of the later, more successful settlements along the upper Paria River (Figure 4). Not all settlements surrounding the Monument were established under Church direction, although new towns still tended to follow the general Zion town plan. Non-mission driven settlement occurred in episodes near the northern end of the Monument in Escalante, Salt Gulch, and Boulder; and in Bryce Valley at Tropic, Cannonville, and Henrieville (and the now-abandoned communities of Loosee, Woodenshoe, and Georgetown). Within the Monument region specifically, the most important mission was sheep and cattle grazing, along with dairy operations. And with increasing human pressures, lands became degraded by unregulated grazing – a fact that would become apparent by the 1930s. As Mormon settlements grew, the way-of-life for Native Americans was unalterably changed. In a few short years, livestock grazing by settlers reduced native grasses and other plants essential to the foraging customs of the tribes.

Photo credit: BLM. Figure 4. Watson cabin, GSENM.

Complimentary Contributor Copy

Land As Sustenance and Sanctuary

185

Game and other resources were depleted or cut-off by settlers, and the Region’s limited waterways were dammed and rerouted, with water allocated to farms and ranches. While Euro-Americans saw their taming of the desert as progress, the tribes saw nothing but devastation and limited access to resources. These cultural incompatibilities arose from radically different paradigms about ownership, land use, and how humans should interrelate with the landscape. The Blackhawk War (1865-1872), fueled by Ute and Paiute frustrations, led to the temporary abandonment of many Mormon settlements from Kanab through Long Valley. Eventually, the Mormons were able to initiate peace and many towns were reoccupied while new ones were settled. In 1879, a group of Saints were called to settle along the San Juan River in southeast Utah via a route across the current Monument. The Hole in the Rock Expedition (1879) has since become legendary for its struggle through “the most unpromising route across some of the roughest country in the United States” (Topping 1997), building a wagon route through sheer cliffs to cross the Colorado River – a feat of tenacity, ingenuity, hard work, and faith. The GSENM region also witnessed the peak of one Mormon social experiment – the United Order – a utopian concept of collaboration and communal ownership patterned after a model introduced by LDS founder Joseph Smith. Although most Orders failed within a few years, the concept reached its pinnacle in a decade-long communal effort in Orderville in the 1870s.

John Wesley Powell and Scientific Exploration In 1866, Captain James Andrus led the Mormon cavalry of the Nauvoo Legions through the present-day GSENM and reached the banks of the Dirty Devil River. A few years later, in 1869, John Wesley Powell (1834-1902) initiated his legendary exploration of the Colorado River. Later, Powell successfully lobbied Congress to fund a second expedition in 1871. This expedition expanded its focus to the Colorado Plateau watershed and began charting, in earnest, the last unmapped and most remote region of the continental United States. The expedition also described the region’s flora, fauna, and geology, forever altering the view of Utah from the outside. The significance of Powell’s Expeditions cannot be overstated. His efforts resulted in the establishment of baseline and triangulation points for the survey and mapping of the interior of the Colorado Plateau, and set the scientific standard for the time (Figure 5). Expedition members were the first Euro-Americans to reach the confluence of the Dirty Devil and Escalante Rivers with the Colorado River, and locate the remote Henry Mountains – the last mountain range discovered in the “Lower 48” states. Unlike many of his contemporaries, Powell understood the limitations imposed by the region’s aridity, and his 1878 report to Congress recommended public land disposal and irrigation projects (Powell 1878). Powell drew upon his experience in Utah to promote a more democratic social organization where groups of families would protect communities and create “irrigation districts” to avoid the control of land and water by a handful of powerful elites. In 1879, Powell became the first Director of the new Bureau of Ethnology. To Powell, Manifest Destiny demanded that Euro-Americans would prevail over the West’s tribal cultures. His views – widely shared at the time – led him to document tribal life as a way to preserve aspects of cultures destined to be lost, although he sometimes staged portraits to support his lobbying efforts.

Complimentary Contributor Copy

186

Robert J. Lilieholm and Marietta Eaton

Photo credit: Robert J. Lilieholm. Figure 5. Powell Point on the Aquarius Plateau, north of the GSENM.

Complimentary Contributor Copy

Land As Sustenance and Sanctuary

187

In 1881, Powell became Director of the United States Geological Survey. At the time, his scientific discoveries drew increased attention to the region, spurring the outside world to increasingly encroach upon the Mormon sanctuary of southern Utah. His vision continues to influence the region today.

Railroads and Non-Mormon Settlers Enter the Region On the surface, Mormon dominance of the region remained strong, but the growing influence of non-LDS or Gentile culture became harder to ignore. Thousands of non-Mormon settlers headed for California during the Gold Rush of 1848-1855, and even more arrived with the completion of the transcontinental railroad, which passed just north of Salt Lake City in 1869. The arrival of new settlers seeking their fortunes in the West was unstoppable, and with them came a host of competing values. The communities surrounding GSENM remained insulated in terms of their world view, sense of community, and spirit. But all around them, resources were increasingly being used by these “newcomers.” Indeed, the region’s growth provided economic opportunities that proved irresistible to Mormon and non-Mormon alike, and ranches sprang-up in remote locations. New people with new ideas arrived, and the isolation that had for a while protected the Saints was no longer a deterrent to the outside world. The area was now well mapped so that obstacles to progress could be avoided. Mormon polygamists came under increasing attack for their practice of plural marriage. In 1882, “The Raid,” initiated under the Edmunds-Tucker Act, included economic, social, and political pressures in its prohibition of polygamy, fragmenting communities and turning many prominent LDS leaders into fugitives. Today, rural communities in southern Utah remain connected to the concept of the frontier and their significance as a peculiar, persecuted people. It has been this peculiarity that has allowed them to retain the core cultural identity that many of their ancestors worked so hard to establish.

Resource Depletion and the Rise of the Conservation Movement Commercial livestock operations began in earnest in the GSENM region in the late 1860s. Shortly thereafter, a severe drought in California forced the movement of large flocks of sheep to the Arizona Strip and beyond – further stressing the productivity of the land. Unregulated livestock grazing degraded range conditions as early as 1890. Outside interests moved into the few areas with remaining forage, intensifying the crisis. In 1894, a six-year drought brought catastrophic results and led to the out-migration of both people and livestock. A classic “Tragedy of the Commons” (Hardin 1968) was unfolding, wreaking havoc on the region’s arid rangelands and ultimately compromising everyone’s welfare. Meanwhile, new fences restricted communal access to rangelands, and new ranchers arrived that could afford to pay the fees that began to be charged for grazing on public lands. As a result, many locals were displaced from grazing on the newly-created Forest Reserves (today’s National Forests) and National Monuments, although not on unregulated Department of Interior (DOI) lands, such as those that eventually became GSENM.

Complimentary Contributor Copy

188

Robert J. Lilieholm and Marietta Eaton

The onset of World War I spurred huge demands for meat, and in 1917 the government permitted more livestock across the entire region, attracting outside operators as well. As noted by Gregory and Moore (1931), “[o]verstocking of the range in response to the increased value of cattle during the World War appears to have been the first step toward the present unfortunate state.” The end of the War led to a crash in livestock markets, exacerbated a decade later by the Great Depression. Yet land degradation continued. Hardship, self-sufficiency, and tenacity remained the hallmarks of the prevailing Mormon culture, but by the 1930s changing values were emerging: “It [once] was a vast blooming flower garden as far as one could see. Now tumble weeds, Russian thistle has taken the place of those hundreds of square miles of country, once so lovely with its covering of luscious grass and other rich forage. Sheep and drought have been the transformers that have changed a veritable Paradise into a desert waste.” (Ott n.d.)

In 1934, the DOI estimated that 155,000 cattle and 250,000 sheep from nearly 19,000 farms and ranches used southern Utah and northern Arizona (Arrington 1986, Hinton 1986). That same year the Taylor Grazing Act was passed in an effort to “stop injury to the public grazing lands by preventing overgrazing and soil deterioration, to provide for their orderly use, improvement, and development, [and] to stabilize the livestock industry dependent upon the public range” (Muhn and Stuart 1988). The Act created the Division of Grazing and directed states to divide lands into districts to be managed by local grazing advisory boards. The situation pitted small operators against large, and with the implementation of the “commensurate property” rule – which required livestock permittees to own base property with water for use if livestock could not graze on public lands for a period of time – forced more small ranchers off the range (Raymond 2003). As a result, many in the region lost their livelihood and the life of the cowboy changed forever – “the result that competition, individualism and private conquest ascended with the rise of secular authority” (Cassidy and Truman n.d.). Communities became less immune to the outside world, and the twin encroachment of market forces and government regulation eroded the region’s insularity and independence. No longer was self-sufficiency and isolation protection from the social and economic depression of the Nation. These economic trends forced many to leave their communities in order to make a living elsewhere. By 1935, 70% of Escalante’s 1,000 residents were on government relief (Cannon 1986). The Civilian Conservation Corps (CCC) began work in the region in 1933, with most efforts focused on range improvements, local sanitation projects, roads, and bridges. Today, remnants from their labors dot the GSENM landscape – scattered corrals, residential pit latrines, extended phone lines, a supply house in Kitchen Corral, and the noteworthy road and bridge over Hell’s Backbone that connected the communities of Escalante and Boulder in 1935 (Woolsey n.d.). In fact, the current route of Highway 12 between these two towns was completed by the CCC in 1940, with the road to Loa finished shortly thereafter. Only more lasting an influence was perhaps the large influx of CCC workers that eventually married local women and stayed in the area to raise new families.

Complimentary Contributor Copy

Land As Sustenance and Sanctuary

189

The CCC mirrored other, more grandiose federal efforts to promote social welfare – FDR’s “New Deal” of rural electrification, federally insured bank deposits, worker safeguards and protections, relief programs, and Social Security. But Pearl Harbor and the onset of World War II lead to the disbandment of the CCC’s workforce, and many projects were abandoned. WWII exacerbated the social and economic isolation of southern Utah, as men and women joined the war effort in numbers sufficient to change the character of their communities. At War’s end, there would be no return to a simpler past. Indeed, the political and economic operations of the region were increasingly integrated into larger market systems under authorities separate from the Church. In 1946 the Grazing Service was combined with the General Land Office to form the Bureau of Land Management (BLM). The new agency was highly decentralized and adopted a “multiple use” mandate under its new Director, Marion Clawson. In 1964 the BLM “reorganized to reflect new programs and authorities under this mandate: concerns for wildlife, recreation, soil, and water resources were integrated into traditional programs (range, forestry, lands, and minerals) through a land use planning process.” (Muhn and Stuart 1988)

Land use planning on BLM lands was finally institutionalized, resulting in more comprehensive and large scale planning – an approach that threatened grazing because other oftentimes competing uses were now given formal consideration in the planning process. Hence, the BLM’s broadened mission, driven by an increasingly national constituency, further limited economic options for the region’s traditional lifestyles.

Literature, Movies, and the Rise of Romanticism In 1910 Zane Grey published The Heritage of the Desert, followed in 1912 with the legendary Riders of the Purple Sage. Between 1914 and 1928, Grey published 17 novels describing the social changes underway in the West. In Wild Horse Mesa (1928), Grey used the Monument’s Kaiparowits Plateau as a backdrop to his idealized view of the region. The mythical Cowboy was further preserved by the illusion created by Tom Mix in a 1922 movie filmed in Kanab, Utah – Deadwood Coach. Using Kanab as a base of operations, film-making became a bonafide industry in this small community until the 1960s. Movie’s filmed in the Kanab area read like a list of who’s who in Westerns from the 1940s through the 1950s: John Ford and Cecil B. DeMille, Union Pacific and Stagecoach (1939), Billy the Kid (1941), In Old Oklahoma (1943), Buffalo Bill (1944), The Green Grass of Wyoming (1948), Fort Apache (1948), She Wore a Yellow Ribbon (1949), Cattle Drive (1950), Westward the Women (1952), Ride, Vaquero (1953), Fort Yuma (1955), Jubal Troop (1956), War Drums (1957), and Fort Bowie (1959), among others. The decades-long film industry exerted a strong influence on the local economy. Although film jobs were sporadic, many locals worked for different productions over the years, and Kanab’s service industry flourished as film crews spent freely. Hollywood stars were themselves subjects of much interest, and the cult of celebrity further inspired the tourism industry. The ancillary combination of movie stars and tourists further changed the face of local communities. In fact, it is interesting to note that this era pitted the “real

Complimentary Contributor Copy

190

Robert J. Lilieholm and Marietta Eaton

cowboys” of the past with the “reel Cowboys” depicted in movies. But the fantasies of these innocent entertainments forever altered the national and even local perception of cowboys and the Western lifestyle (Figure 6). An archetype had been born. The appeal of Westerns came largely from straight forward and simple plots – a more often than not inaccurate depiction of reality. Within this narrative, the Cowboy came to represent the upstanding and simple virtues of Western pioneers. Indeed, the individualistic heroes of the screen were a far cry from the focus on communal cooperation, hard work, and family that prevailed in the region. The resolution of conflict on film by gun had little basis in fact. Nonetheless, the austere landscape of GSENM provided a romantic palette upon which to paint a wild, untamable, and rugged picture of the past – an escape to what for many was their only perspective of life in the West.

Photo credit: BLM. Figure 6. Ranching remains an important way of life for local residents.

The advent of television forever changed the face of entertainment, and several TV series were filmed in the area, including The Lone Ranger, Death Valley Days, Have Gun Will

Complimentary Contributor Copy

Land As Sustenance and Sanctuary

191

Travel, Wagon Train, Cheyenne, and Gunsmoke. Movies were still filmed in and around Kanab during the 1960s and 1970s, with such fare as Sergeants Three (1963), MacKenna’s Gold (1969), The Outlaw Josie Wales (1976), and Disney’s The Apple Dumpling Gang Rides Again (1978). But as interest in Westerns faded, the movie industry left Kanab, although a few non-Westerns were filmed there near the end of the heyday of “Little Hollywood,” including The Greatest Story Ever Told (1969), Planet of the Apes (1968), Exorcist II, and The Heretic (1977). The last movie of any substance was Windrunner in 1993.

Economic Integration and the Rise of Tourism While grazing and farming continued to be important for many communities, mining, logging, and tourism also emerged as economic forces. Mining was sporadic and seldom profitable. Coal was mined for local use, and prospecting for gold and copper had occurred in the 1880s. In 1915, large coal deposits and likely areas of oil, gas, and tar sands were discovered near the Circle Cliffs (Gregory and Moore 1931). During the Cold War, the Atomic Energy Commission sought reliable sources of uranium. Again, the Circle Cliffs experienced a boom that lasted until the mid-1950s. The quality of the resource proved marginal, but the traces of this era remain today, with the scenic Burr Trail being most notable. In the 1970s, Kaiparowits coal was slated for future mining until the designation of the Monument. Tourism has played an important economic and social role in the region for over a century. Americans began visiting the Grand Canyon in the 1890s – about the same time it received its first protection as a Forest Reserve. Other notable tourist destinations near GSENM include Mukuntuweap National Monument protected in 1909 (today’s Zion National Park), Grand Canyon National Park (1906), Utah National Monument (1923) (today’s Bryce Canyon National Park), Pipe Spring National Monument (1923), and Cedar Breaks National Monument (1933). In 1928, Harvard Professor Clyde Kluckhohn visited the largely-unexplored Kaiparowits Plateau and lobbied for creation of a national park – that is until he realized that designation would alter the very isolation that made the region so special. National lobbying efforts by rail, auto, and highway interests continued, however, further eroding the remoteness of southern Utah – especially once a bridge across the Colorado River opened near Lee’s Ferry in 1929. While tourism remained a suspect industry that catered to “outsiders,” local entrepreneurs increasingly stepped in to accommodate visitors. Ruby’s Inn near Bryce Canyon National Park has been a popular tourist destination since 1916 and today is the largest employer in Garfield County. The Parry brothers of Kanab built a lucrative tourist business providing lodging and bus tours to Zion, Red Canyon, Bryce Canyon, and the North Rim of the Grand Canyon. In 1936 the National Park Service (NPS) considered the creation of a new national monument that would include much of the Escalante River drainage. Then as now, locals rose in opposition, fearing the loss of grazing lands (Richardson 1965). Support for protection was limited, and the plan languished and died in 1940 as national attention turned toward the growing conflict in Europe. A 1947 motorized “expedition” into the GSENM area by the National Geographic Society further exposed the region to a national audience (Breed 1949).

Complimentary Contributor Copy

192

Robert J. Lilieholm and Marietta Eaton

Nearby areas were also being considered for resource protection. Capitol Reef National Monument, along GSENM’s eastern boundary, was established in 1937 despite intense opposition from local ranchers (Frye 1998). The history of Capitol Reef is rife with conflict, most of which centered on the acquisition of private lands in the nearby community of Fruita. There, the town witnessed an influx of tourists and “outsiders” after the Monument was created, and by the late 1950s was no longer insulated from the outside world. But the town’s limited tourist facilities lead the government to begin acquiring private lands – a controversial act that fueled local frustrations. Today, local legend holds that this was yet another federal effort to prevail over rural Utahans – part of a relentless persecution that continues. Fear of a similar fate alarmed many GSENM communities following the Monument’s designation in 1996. In 1964, the completion of Glen Canyon Dam on the Colorado River had a profound influence on Monument communities. Glen Canyon City (now Big Water) emerged from the desert on the Utah side of the River, a boom-town arising from the Bureau of Reclamation’s mission to not only provide water and electricity, but foster recreation and economic development as well. The dam completed access to the Colorado River, beckoning more tourists and “outsiders.” Such rapid development strained the infrastructure of Kane County and drew an ever-rising influx of tourists. Today, the 1.25-million-acre Glen Canyon National Recreation Area draws more than 1.7 million visitors each year – over 300 tourists for every County resident. In the 1980s and 1990s, hostility over the designation of Forest Service lands for inclusion in the National Wilderness Preservation System became yet another flashpoint for communities in southern Utah. Seen as federal-level interference fueled by environmentalists and “Eastern elites,” many had believed that the limited amount of Forest Service lands set aside in the 1980s had settled the dispute. Locals felt betrayed when additional Forest Service lands were considered, and then when the BLM began its own round of wilderness reviews in what seemed like a never-ending process.

Designation of the GSENM On September 18, 1996, President Bill Clinton established the GSENM under the authority of the 1906 Antiquities Act. Although the official impetus was to protect the 1.96 million-acre region’s geological, paleontological, archeological, historical, and biological resources, the creation of the Monument in many ways settled a decades-old debate over wilderness designation in the State. Utah’s intransigence over the wilderness issue – and Clinton’s somewhat heavy-handed response – was not unanticipated. As noted by Lilieholm (1995) a year before GSENM’s creation: “… the federal government prefers to defer to states when federal resources are threatened. However, the federal government will assume [protection] responsibility if states are either unwilling or unable to act… Federal deference to state rule has important implications for the wilderness debate in Utah. For example, state cooperation in wilderness management and protection reduces the likelihood of federal involvement. State and local hostility toward wilderness may invite an expanded federal role in

Complimentary Contributor Copy

Land As Sustenance and Sanctuary

193

protecting wilderness resources, which could ultimately compromise both state and federal interests in resource protection.”

The ensuing federal action and state response were swift. Locals complained that the planning for the Monument took place in secret without input from Utah’s political leadership – a view not shared by all (see, for example, Leshy 1998). Adding insult to injury, the President’s signing ceremony took place across the state line in Arizona. Some suggested that the region’s protection served to garner environmental votes in the upcoming presidential election, with little political cost in the way of lost votes from the Republican stronghold of Utah (Malakoff 1999). Utah’s Congressional Delegation ordered investigations into the designation process, and Kane and Garfield Counties sent bulldozers into the Monument to groom roads in an effort to ensure continued access (Rusnak 2003).

LESSONS LEARNED AND FUTURE DIRECTIONS Social Transformation and the Struggle over Local Versus National Control Over the last century, the mission of many public lands in the US has evolved from one of commodity-driven exploitation, to one centered around aesthetic enjoyment and the provision of ecosystem services (Brunson and Steel 1994). The result is that by the latter half of the 20th Century – just 100 years after Mormon settlers first arrived in the GSENM region – southern Utah’s social fabric had been totally altered, first by overuse of the land, then regulation of the public domain, and finally by a new focus on recreation, tourism, and ecosystem health and restoration. The changing focus of public land management was not the only challenge. Rural communities struggled as the post-WWII national economy displaced local production systems, and as American culture was homogenized through national markets, mass marketing, and television. Economic integration and globalization, beginning in the 1980s, further strained local economies, and the rise of the automobile, highways, and telecommunications eroded the isolation of once-remote communities around the Monument. For many locals, the Monument’s designation in 1996 was the embodiment and capstone of these trends, rekindling a host of long-simmering fears about past persecution, outside influence, and the loss of access to traditional resources. Indeed, the relentless push of national-level demands for protected landscapes and recreational opportunities challenged the core identity of Monument communities, affecting them economically, culturally, and spiritually. Most vexing was the realization that many of these changes were driven by a national – and even international – environmental constituency. Within such a highly charged political environment, it is all too easy for land managers to forget that historical relationships matter when it comes to understanding how local communities interpret and react to change. The intensity of feelings pent-up from more than a century of mistrust can surprise bureaucrats caught unaware of the social and historical context within which policies are developed and lands managed. For GSENM, this past presents unique challenges for future management. Indeed, as the largest US reserve ever created specifically for scientific purposes (Malakoff 1999) – and the first National Monument to be placed under BLM jurisdiction – there are few examples to guide planning and management.

Complimentary Contributor Copy

194

Robert J. Lilieholm and Marietta Eaton

Maintaining Traditional Lifestyles and Working Landscapes In a 1997 speech, Bruce Babbitt, President Clinton’s Secretary of the Interior, referred to GSENM as a “different kind of place.” Babbitt understood the need to pursue a new approach to managing federal lands, and as a Westerner he understood the role of local communities in resource stewardship, as well as the complex and historic interrelationships between land and people. Babbitt noted that “these landscapes are part and parcel of the history, and the context, and the culture of the communities that are placed thereon. In this case, we are administering the assets in the context of one of the richest historic traditions that has ever grown up in the United States or anywhere else. The people in this region do live lightly on the land. This is evident to anyone who has ever on a summer day approached a southern Utah town on a rural road. You first see the Lombardy poplars on the horizon, those wonderful straight lines into a community or village. You see wide roads with irrigation ditches bubbling. You see a town surrounded by agricultural fields of a human scale being tended by one family at a time. That landscape exists in the context of a community whose institutions and whose values and beliefs have been conditioned deeply by their history, their leadership, and their love and their relationship to that landscape.”

Battle lines between use and preservation were apparent soon after designation of the Monument, and questions grew as to how the BLM would implement its multiple-use mandate. Even with a diverse planning team, fears lingered. Perhaps the clearest expressions of local frustration were found in debates over proposed federal restrictions on motorized access and grazing practices (Malakoff 1999) – proposals that struck at the heart of the local culture by questioning the future viability of traditional livelihoods (Brugger 2008). Indeed, grazing was practiced to some degree by nearly every pioneer family, and although diminished by time, it is still this vision that remains the “tradition” of the region (Eisenhauer et al. 2000) (Figure 7). Ranching is rooted in the desire for an enduring connection to the land, the craving for isolation in a society where it is less and less easy to find, the genuine concern for raising a family in a simpler environment, and the strong sense of ownership of nearby public lands. By carrying on a multi-generational lifestyle, these values are passed on and shared. The tradition of family heritage, religious conviction, and connection to the past are important ideals to embrace in an ever-changing world. The transition of public lands toward non-commodity uses presented another challenge as “newcomers” were drawn to newly-protected landscapes and budding “gateway” communities (Figure 8). Some came to visit; others stayed, perhaps starting new businesses or their retirement years. In many locales, these changes created both opportunities (e.g., inflow of new people and ideas, capital and expertise, rising property values) and challenges (loss of control and decreased affordability), but for many residents, the perceived costs outweighed the gains. And ironically, while traditional livelihoods became increasingly marginal in an economic sense, they remained a core element in defining the mystique and attraction of the land for both residents and visitors alike.

Complimentary Contributor Copy

Land As Sustenance and Sanctuary

Photo credit: Robert J. Lilieholm. Figure 7. Cannonville, Utah.

Photo credit: Robert J. Lilieholm. Figure 8. BLM’s Cannonville Visitor Center, GSENM.

Complimentary Contributor Copy

195

196

Robert J. Lilieholm and Marietta Eaton

Land As Sustenance and Sanctuary Amid these changes, Monument communities struggled to maintain their traditions and institutions in the face of incredible odds by turning inward for guidance. Even today, Mormon society remains close-knit and connected to a vision that highlights religious and family values and hard work, with more of an interest in raising kids than cows. Caught between the many facets of this battle is the long-standing desire of prevailing Mormon society to shield itself from unwanted change, to adhere to the tenets of the Church, and to honor the past by preserving local traditions. These competing interests place residents in a cultural vice from which there is no easy extrication – a view likely to intensify as LDS culture becomes increasingly distinct from the broader US mainstream (Toney et al. 2003). Throughout southern Utah’s settlement, however, there is one common theme that cuts across both time and cultures – the importance of the landscape for both sustenance and sanctuary. For the GSENM today, the land itself and the isolation it offers have become its most valuable asset in the eyes of both residents and visitors. This feature may not last long. The Colorado Plateau is surrounded by some of the most rapidly growing metropolitan areas in the US, and out from these urbanized settings file people looking for the peace and solitude that the Monument has always offered in abundance. As “civilization” has progressed, so too has our desire – for both Mormons and non-Mormons alike – to seek refuge (Figure 9). And these pressures will likely intensify as the region becomes more of an international destination.

Photo credit: Robert J. Lilieholm. Figure 9. Recreationist explores one of GSENM’s many slot canyons.

Complimentary Contributor Copy

Land As Sustenance and Sanctuary

197

Photo credit: Robert J. Lilieholm. Figure 10. View from Cottonwood Road near Kodachrome State Park, GSENM.

Already, nearly one-quarter of GSENM visitors are from overseas, compared to less than 4% of tourists statewide (Burr et al. 2006). Preserving the region’s solitude in the face of rising visitation will likely be a central feature in future debates. Television, highways, air travel, iPhones and the internet have changed our lives forever, yet the isolation that many seek is only to be found by exploring the far reaches of even this remote place (Figure 10). More so than anywhere else in the “Lower 48,” the GSENM is still a place where one can retreat from our hectic world. In fact, it is this very thing that is priceless and endures, and what people past and present have come here to find.

ACKNOWLEDGMENTS This research was supported by National Science Foundation award EPS-0904155 to Maine EPSCoR Sustainability Solutions Initiative at the University of Maine. Additional support was provided by the Maine Agricultural and Forest Experiment Station, and the Center for Research on Sustainable Forests.

Complimentary Contributor Copy

198

Robert J. Lilieholm and Marietta Eaton

REFERENCES Arrington, L. J. 1986. Utah’s great drought of 1934. Utah Historical Quarterly 54:258-259. Breed, J. 1949. First motor sortie into Escalante land. The National Geographic Magazine XCVI: 369-404. Brugger, J. V. 2008. The “other” and “the enemy”: Reflections on fieldwork in Utah. North American Dialogue 10(1):6-10. Brunson, M. W. and B. S. Steel. 1994. National public attitudes toward federal rangeland management. Rangelands 16:77-81. Burr, S. W., D. J. Blahna, D. Reiter, E. C. Leary, and N. M. Wagoner. 2006. A Front Country Visitor Study for Grand Staircase-Escalante National Monument. Institute for Outdoor Recreation and Tourism, Utah State University. IORT Professional Rep. PR2006-1. 155 pages. Cannon, B. Q. 1986. Struggle against great odds: Challenges in Utah’s marginal agricultural areas, 1925-39. Utah Historical Quarterly 54(4):314-315. Cassidy, M. and K. Truman. n.d. Grand Staircase-Escalante National Monument historic resources overview (manuscript on file, GSENM Headquarters Office, Kanab, Utah). Eisenhauer, B. W., R. S. Krannich and D. J. Blahna. 2000. Attachments to special places on public lands: An analysis of activities, reason for attachments, and community connections. Society and Natural Resources 13:421-441. Foltz, R. C. 2000. Mormon values and the Utah environment. Worldviews 4:1-19. Frye, B. J. 1998. From barrier to crossroads: An administrative history of Capitol Reef National Park, Utah. USDI National Park Service, Intermountain Region, Cultural Resources Selections, No. 12. Gregory, H. E. and R. C. Moore. 1931. The Kaiparowits region: A geographic and geologic reconnaissance of parts of Utah and Arizona. Government Printing Office, Washington, DC. Hardin, G. 1968. The tragedy of the commons. Science 162:1243-1248. Hinton, W. K. 1986. The economics of ambivalence: Utah’s Depression experience. Utah Historical Quarterly 54(3):278-287. Jackson, R. H. 1978. Mormon perception and settlement. Annals of the Association of American Geographers 68(9):317-334. Leshy, J. D. 1998. Putting the Antiquities Act in perspective. Pages 83-88 In: R. B. Keiter, S. B. George and J. Walker, eds., Visions of the Grand Staircase-Escalante: Examining Utah’s newest National Monument. Salt Lake City, UT: Utah Museum of Natural History and Wallace Stegner Center. Lilieholm, R. J. 1995. Wilderness and adjacent land issues. Chapter 6, In: D. L. Snyder, C. Fawson, E. B. Godfrey, J. E. Keith, and R. J. Lilieholm, eds. Wilderness Designation in Utah: Issues and Potential Economic Impacts. Utah Agricultural Experiment Station Research Report 151. Malakoff, D. 1999. Proposed access rules split community. Science 283:1619. Muhn, J. and H. R. Stuart. 1988. Opportunity and challenge: The story of BLM. Washington, DC: Government Printing Office. Ott, L. J. n.d. Life sketch of Joseph Wallace Thompson,” Federal Writers’ Project, WPA Papers, Utah State Historical Society Collections, Salt Lake City.

Complimentary Contributor Copy

Land As Sustenance and Sanctuary

199

Powell, J. W. 1878. Report on the lands of the arid region of the United States, with a more detailed account of the lands of Utah. Boston, MA: The Harvard Common Press (1983 reprint). Raymond, L. 2003. Private rights in public resources: Equity and property allocation in market-based environmental policy. Washington, DC: Resources for the Future Press. Richardson, E. R. 1965. Federal park policy in Utah: The Escalante National Monument controversy of 1935-1940. Utah Historical Quarterly 33:109-133. Rusnak, E. C. 2003. The straw that broke the camel’s back? Grand Staircase-Escalante National Monument antiquates the Antiquities Act. Ohio State Law Journal 64:669-730. Toney, M. B., C. Keller and L. M. Hunter. 2003. Regional cultures, persistence and change: A case study of the Mormon Culture Region. The Social Science Journal 40:431-445. Topping, G. 1997. Glen Canyon and the San Juan Country. Moscow, ID: University of Idaho Press. Woolsey, N. n.d. Life in Boulder, Utah: The King story. Scottsdale, AZ: Agreka Books.

Complimentary Contributor Copy

Complimentary Contributor Copy

In: National Parks Editor: Johnson B. Smith

ISBN: 978-1-62618-934-8 © 2013 Nova Science Publishers, Inc.

Chapter 7

ASSESSING THE COMPETITIVENESS OF PROTECTED AREAS: THE COMPETITIVE POSITIONING OF THE PENEDA-GERÊS NATIONAL PARK Maria João Carneiro* and Carlos Costa† Research Unit in Governance, Competitiveness and Public Policy of the University of Aveiro, Portugal Department of Economics, Management and Industrial Engineering of the University of Aveiro, Portugal

ABSTRACT Protected areas stand out due to the importance of specific nature and biodiversity features they comprise. However, several researchers highlight that the competitiveness and consequent sustainability of tourism destinations do not only depend on a specific resource of the destination, but on the whole basis of resources of the destination and on the way this set of resources is managed. Competitive positioning studies became an important tool for evaluating the competitiveness of destinations, since they enable to understand how potential visitors evaluate the destination comparing to potential competitors. Therefore, this kind of studies provides important contributions for improving the management and competitiveness of destinations. Despite the importance of this type of research, positioning studies on protected areas are rare. The present chapter aims to: highlight the importance of competitive positioning studies of protected areas; present a methodology that may be used for assessing the competitive positioning of protected areas; and to assess the competitive positioning of a specific national park located in Portugal - the Gerês National Park. It is also aimed to provide contributions for a more efficient management of the Gerês Park, in particular, and of protected areas, in general. A total of 1115 questionnaires directed to visitors of Gerês were obtained. Results highlight the importance of assessing the positioning of protected areas considering the performance of these areas regarding the following features: tourism attractions; * †

E-mail: [email protected]. E-mail: [email protected].

Complimentary Contributor Copy

202

Maria João Carneiro and Carlos Costa facilities; protected areas’ ability to satisfy motivations; and the constraints visitors feel for visiting these areas. The empirical research undertaken reveals that Gerês holds competitive advantages in relation to competing destinations in several domains, such as on natural attractions, peacefulness and on providing opportunities for escaping and relaxing. However, the study also highlights the need to further explore the potential of protected areas, by complementing the important natural supply of these areas with, for example, an attractive and well managed set of cultural resources and interesting opportunities for socializing. The provision of information about the protected area reveals to be a good strategy for improving the attractiveness of protected areas, as well as for reducing constraints for visiting these areas.

1. INTRODUCTION Kotler (1997) stresses the importance of designing the supply by adopting approaches that enable companies to achieve a competitive position in the minds of consumer. This position must be valued by consumers and distinct from the positions of competitors. Dwyer and Kim (2003) also postulate, that to achieve a competitive advantage, destinations should have an attractiveness level and provide experiences that are superior to those of competing destinations. For being competitive, destinations must than have a better performance than the competitors and this difference of performances must be visible and valued by consumers. In the last decades, there has been a growing interest on the evaluation of the competitiveness of tourism destinations based on visitors’ perspectives. However, most of the studies on positioning of destinations do not encompass a broad range of the factors that can influence the positioning of destinations. Moreover, the majority of the studies regarding the positioning of tourism destinations relate to countries (e.g. Baloglu and McCleary, 1999; Haahti, 1986; Kim et al., 2005; Kozak et al., 2010), and cities (e.g. Dolnicar et al., 2000; Woodside and Lysonski, 1989; Jin et al., 2010). The assessment of the competitiveness of protected areas has been widely neglected. This chapter extends research in the field of the positioning of destinations by suggesting a comprehensive methodology to analyze the competitiveness of protected areas and by providing an assessment of the competitive positioning of the Gerês National Park - a protected area located in Portugal. Apart from the present introduction, the chapter includes a section of literature review, a section related to the empirical study and, finally, a section referring to conclusions and implications. The literature review section begins with a literature review on competitiveness and positioning studies of destinations and continues with a literature review on the positioning of protected areas. The section about the empirical research, first focuses on the methodology used for developing the empirical research and, then, on the presentation and discussion of the results of the empirical study, concerning the positioning of the Gerês park. The chapter ends with the presentation of conclusions and implications derived from the literature review and from the empirical research. In this last section, considerations are also done regarding some limitations of the of the study and suggestions for further research.

Complimentary Contributor Copy

Assessing the Competitiveness of Protected Areas

203

2. LITERATURE REVIEW 2.1. Evaluation of the Competitiveness of Tourism Destinations Tourism attractions have a very relevant role in the competitiveness of tourism destinations since they are the components of the tourism destinations that have the power of attracting visitors to the destinations. Attractions, also designed as pull factors (Crompton, 1979) correspond to components that, individually or in group, motivate the visit to the destination (Middleton, 1989). Both models of destination competitiveness proposed by Ritchie and Crouch (2003) and Dwyer and Kim (2003) remark the main role of resources and attractions as core elements of destinations. However, in order to ensure that visitors will be able to visit and enjoy the attractions, destinations have also to provide appropriate facilities to support tourism (Middleton and Clarke, 2001; McIntosh et al., 1995). Inskeep (1991), in his research on tourism planning, provides a useful insight on the range of tourism attractions and of facilities and services to support tourism. Echtner and Ritchie (1993) also suggest a comprehensive list of tourism attractions as well as of facilities and services that should be considered in studies for evaluating the image of tourism destinations. Tourism attractions identified by these researchers comprise attractions so distinct such as flora, fauna, beaches, museums, architecture and buildings, historic sites, and gastronomy, among others. As far as facilities and services to support tourism are concerned, the research above referred reveals that accommodation and accessibility infrastructures are of special importance for destinations. However, they also highlight that these facilities and services also have to be complemented by other important infrastructures such as those related to the provision of touristic information or to the provision of food and beverage (such as restaurants). Both Ritchie and Crouch (2003) and Dwyer and Kim (2003) advocate that, despites the relevance of tourism attractions and supporting facilities and services, the competitiveness of a tourism destination also highly depends on the way destinations are managed. Taking into consideration the importance of the management of destinations, it is interesting to notice that the model of Dwyer and Kim (2003) highlights that the competitiveness of destinations is also determined by situational conditions. Furthermore, destination choice models proposed by several authors (e.g. Mill and Morrison, 1998; Moutinho, 1987; Orth and Turecková, 2002; Ryan, 1994; Um and Crompton, 1990; Woodside and Lysonski, 1989) suggest that motivations and structural constraints are likely to influence the decisions of visitors when selecting a tourism destination to visit. Motivations are states or driving forces that impel people to engage in a certain behavior, to carry out actions, in order to reduce a tension that they feel (Moutinho, 1987). Several researchers (Beard and Ragheb, 1983; Crompton, 1979; Lee and Crompton, 1992; Manfredo et al. 1996) developed research in order to identify a range of leisure or tourism motivations. Leisure and tourism motivations identified by these authors and that seem to have a more prominent role in the majority of tourism studies, are associated to features such as novelty, relaxation, escape, socialization and expanding knowledge. Freedom, happiness, competence, regression (e.g. sensation of temporarily regress to early stages of the lifecycle), discovery of the self and prestige are other motivations found that may also assume importance in several scopes of the tourism field.

Complimentary Contributor Copy

204

Maria João Carneiro and Carlos Costa

Tourism constraints are factors that prevent people interested in participating in tourism from participating in it, but that may also influence preferences regarding this participation (e.g. preferences for tourism destinations) (Crawford and Godbey, 1987; Jackson and Scott, 1999). Researchers proposed several typologies of classification of tourism and leisure constraints. Jackson (1993) identifies five leisure constraint dimensions: transportation and access; facilities and opportunities; skills and abilities; costs; and time. However, one of the typologies more widely adopted is that suggested by Crawford and Godbey (1987), which is already adopted in several studies in the field of leisure and tourism (e.g. Pennington-Gray and Kerstetter, 2002; Raymore et al., 1993). According to Crawford and Godbey (1987), constraints may be classified in three categories: intrapersonal, interpersonal or structural. Intrapersonal constraints are “individual psychological states and attributes which interact with leisure preferences” (p.122) (e.g. stress, subjective evaluations of the appropriateness and availability of various leisure activities and anxiety). Interpersonal constraints correspond to barriers emerging as a “result of interpersonal interaction or the relationship between individuals’ characteristics” (p.123) (e.g. lack of company for visiting a destination, having to take care of sons). Structural constraints refer to “intervening factors between leisure preference and participation” (p.124) (e.g. financial resources, time constraints, climate). Structural constraints have a higher prominence than other kind of constraints in destination choice models and, mainly, in literature on positioning of tourism destinations, probably because these constraints are those that are more easily addressed and managed by the managers and planners of tourism destinations. Studies about constraints for participating in leisure and tourism activities (e.g. Tian et al., 1996; Hudson, 2000) also revealed the existence of a wide variety of tourism structural constraints related, among other factors, to money, equipment, time and crowding. The destination choice models above referred suggest that, while motivations are more likely to influence the choice decision at the early stages of this process, the influence of structural constraints is more likely to be stronger at the final stages of the process. Studies on positioning of destinations carried out based on the perceptions of visitors also provide an insight on features that contribute to a higher competitiveness of tourism destinations. In 1986, Haahti (1986) conducted a research to measure the positioning of Finland in relation to other 11 European countries. The assessment is based on 10 items primarily associated with a restricted number of attractions (e.g. nightlife and entertainment, friendly and hospitable people), with facilities (e.g. accessibility/easy to reach) and with the ability to satisfy a reduced number of motivations (e.g. to change from the usual destinations). In 2005, Kim et al. (2005) undertook a study to evaluate the positioning of six countries that Mainland Chinese may consider visiting – Singapore, Thailand, Japan, South Korea, Egypt and Germany. Positioning is identified based on the perceptions of Chinese people mainly concerning the following components: attractions (e.g. good weather, beautiful scenery) and facilities (e.g. good leisure and recreation facilities, well-equipped tourism facilities). Kozak et al. (2010) assess the positioning of Turkey in relation to Greece and Spain on four main components: quality of facilities and activities, cultural and natural attractiveness, quality of services and quality of infrastructure. Cracolici and Nijkamp (2008) compared the attractiveness of Six Southern Italian regions through an index that measures the performance of these regions on attractions (e.g. landscape, environment / nature, cultural events, wine quality) and facilities (e.g. hotels and other accommodation, information and tourist services). As may be observed, most research on positioning of destinations in the perspective of

Complimentary Contributor Copy

Assessing the Competitiveness of Protected Areas

205

visitors relies, mainly, on visitors’ perceptions of destinations regarding attractions, facilities and ability to satisfy some motivations. A limited number of studies on positioning of destinations explicitly encompass the constraints to visit the destination. When assessing the positioning of eight Korean national parks, Hong et al. (2006) consider structural constraints and antecedent constraints (antecedent constraints comprise some intrapersonal and interpersonal constraints). Botha et al. (1999), in a study designed for analyzing the competitive position of Sun/Lost City, in South Africa, considered many of the structural constraints previously referred (e.g. financial constraints such as the trip or the accommodation being too expensive, time constraints, accessibility constraints and constraints related to safety. Taking into consideration the literature reviewed on competitiveness and on positioning of destinations, it is proposed, in this chapter, a methodology for assessing the positioning of protected areas. This methodology corresponds to comparing the protected area with its main competitor in a broad range of tourism attractions, facilities and services for supporting tourism, ability to satisfy motivations and structural constraints to visit the protected area (Figure 1). The objective of this analysis is to analyze whether there are statistical significant differences between the protected area and its competitor and, when these differences exist, to investigate if the protected area has a competitive position in relation to the competitor. This may be accomplished by analyzing if the protected area is more attractive or has a better performance than its competitor. The next section focus on protected areas, and will provide an insight regarding the set of tourism attractions, facilities and services for supporting tourism, ability to satisfy motivations and structural constraints that must be considered when applying this methodology. The complete lists of attractions, facilities and supporting services, motivations and constraints, can be found in the section of the empirical research.

Protected area

Competitor

Attractions

Attractions

Facilities and services for supporting tourism

Facilities and services for supporting tourism

Ability to satisfy motivations

Ability to satisfy motivations

Structural constraints

Structural constraints

Figure 1. Methodology proposed for assessing the competitive position of protected areas.

Complimentary Contributor Copy

206

Maria João Carneiro and Carlos Costa

2.2. Factors Influencing the Competitiveness of Protected Areas Research on the positioning of protected areas, although limited, provides some light regarding factors that may determine the competitiveness of protected areas. Kim (1998) assesses the positioning of five Korean protected areas against each other considering, mainly, attractions, facilities and the ability to satisfy some motivations. This research encompasses attractions such as the natural environment, but also the fresh air, the clean water, the quiet environment, but also other kind of attractions as historic sites and evening entertainment. In the study there is also a reference, among other facilities, to accommodation, food and beverage facilities, and to opportunities for relaxing and for experiencing a new lifestyle. As far as constraints are concerned, this study only includes, in an explicitly way, one situational variable - seasons of the year. The positioning of the protected areas under analysis is evaluated considering the season of the year. Some years later, Deng et al. (2002) conducted a study to analyze the positioning of 36 parks in Australia based on five main components: tourism resources, tourist facilities, accessibility, local communities and peripheral attractions. The research was carried out with the help of a panel of experts, using secondary data about the previously referred components. The Standard Deviation Method suggested the classification of parks into four groups with different levels of attractiveness, and the level of attraction was related to the number of visitors of the parks. In the framework proposed by these authors, there is an interesting identification of several features regarding natural resources (e.g. scientific value, aesthetics and rarity of fauna and flora), but also of cultural resources (e.g. history, architecture and religion). Apart from these resources, the authors also emphasize the role of recreational and educational facilities as well as of accessibility (for example regarding connectivity and accessibility alternatives). In 2006, Hong et al. (2006), as already mentioned, also developed a research in order to investigate the decision-making process of Seoul citizens applied to eight national parks in Korea. In this research, parks are compared on affective images and on a wide range of constraints - antecedent and structural constraints. This study highlights the importance of intrapersonal, interpersonal and structural constraints in the positioning of protected areas (e.g. crowding, nobody to go with, physical effort required, transportation, lack of proper accommodation). In the study of Hsu et al. (2009) the competitive positioning of eight Taiwan touristic spots – three national parks and five other touristic spots – is analyzed. The evaluation and comparison of these eight destinations is focused on the ability for satisfying motivations. Relaxation, novelty seeking, culture exploration, escape and meeting people are some of the motivations considered in the research. However, the study also includes perceptions about a restricted number of attractions and facilities (e.g. transportation facilities, quality of food, environmental safety, friendliness of people and shopping). Other studies apart from those on positioning, also reveal some factors influencing the competitiveness of protected areas. For example, in their research to identify the motivations of ecotourists who visit Anapurna, in Nepal, Holden and Sparrowhawk (2002) identified a wide range of motivations, namely, thrills and excitement, self-esteem and development, social interaction, physical relaxation and fulfillment. When studying the push and pull factors that influence the visit to Korean national parks, Kim et al. (2003) identify several motivations for visiting the parks: being with the family,

Complimentary Contributor Copy

Assessing the Competitiveness of Protected Areas

207

developing friendship, appreciating nature, adventure, having health and getting away from the routine. These researchers also indicate that, for example, rare fauna and flora, environment highly well conserved, tranquil areas, beautiful natural resources, convenient parking and comfortable accommodation, are important pull factors of the protected areas that must be taken into consideration. A survey conducted by Marques et al. (2010) in order to analyze the diversity of domestic visitors to Portuguese protected areas provides a broad range of motivations for visiting protected areas and of facilities and services to support tourism. Enjoying nature, participating in some kind of events, having contact with persons and personal fulfillment, are some of the motivations included in the study. The list of facilities and services to support tourism comprises features such as facilities for providing accommodation and for providing food, interpretation centers, signposts and interpretation facilities (e.g. interpretation centers). As far as constraints are concerned, Marsinko et al. (2002) find that the number of trips people carried out to places that offer wildlife recreation opportunities are influenced by the cost to go on the trip, which includes the time and money required for undertaking the trip. Research on constraints to visit natural resource areas in Michigan (Pennington-Gray and Kerstetter, 2002) and state parks in Pennsylvania (Kerstetter et al., 2002) revealed a wider range of potential structural constraints for visiting protected areas. These studies suggest that time constraints, financial constraints, weather conditions, interpersonal constraints, equipment constraints, overcrowding and lack of skills/ability may be high barriers inhibiting visiting protected areas. Lack of facilities, existence of rules, and safety, although important factors in determining the competitiveness of some tourism destinations not emerge as strong barriers in the studies of Kerstetter et al. (2002) and Pennington-Gray and Kerstetter (2002). There is no consensus between the respondents of the two studies regarding the importance of accessibility and the lack of information. Despites the research undertaken on protected areas, including studies on positioning of protected areas, there is a lack of studies that assess the positioning of protected areas based on a wide range of features representing attractions, facilities, ability to satisfy motivations and structural constraints. The empirical research presented in this chapter attempts to overcome the limitations of the research on positioning of protected areas identified in the literature review of the chapter.

3. COMPETITIVE POSITION OF GERES PARK: EMPIRICAL STUDY 3.1. Methods The aim of the present empirical study is to assess the competitive position of PenedaGerês National Park – a protected area located in North Portugal – against its main competitors. The objective is to evaluate the competitiveness of the Gerês Park regarding facilities and services to support tourism, tourism attractions, ability to satisfy motivations and constraints. This protected area was created in 1971 and is still the only National park in Portugal. It comprises part of the area of five municipalities and has a total area of about 70,000 hectares. It is a mountain area located in the Northwest of Portugal, crossed by several rivers, such as

Complimentary Contributor Copy

208

Maria João Carneiro and Carlos Costa

the Minho, Lima and Cávado rivers. This national park is characterized by a wide variety in terms of fauna and flora highly associated to the geomorphic characteristics of the Park. As far as fauna is concerned, some important species of the park are the trout of the river, the otter and the red-billed magpie (ICNF, 2012). There are also, in the area of the Park, about 15 species of bats and more than 100 species of birds. The municipalities where Gerês is located also encompass important cultural heritage, including 28 national monuments (IGESPAR, 2012). The majority of these monuments is of religious character and is greatly marked by Roman architecture. They include the monasteries of Santa Maria das Júnias and Ermelo, many churches and chapels. Other national monuments of high value are the castles of Melgaço, Montalegre and Lindoso, as well as the bridge over the Lima river.

3.1.1. Survey Instrument In order to accomplish the objectives of this empirical research, it was decided to carry out a questionnaire-based survey directed to visitors of the Gerês park. In order to assess the positioning of the Gerês park in relation to its main competitor, a high number of items related to facilities and services to support tourism, tourism attractions, ability to satisfy motivations and constraints, had to be included in the questionnaire. Therefore, in order to reduce the quantity of items to include in the final questionnaire, it was decided to conduct a pilot test of the list of items with visitors of protected areas in Portugal and with Portuguese university students who had already visited protected areas. After having a complete version of the questionnaire, a pre-test of the questionnaire was undertaken and only minor changes were introduced. The final questionnaire was mainly composed by closed questions and only a very limited number of open questions. Respondents were asked to think about the period before visiting Gerês, when they were planning the trip. They had to indicate the other destinations that they thought about visiting during this period for the purpose of leisure, recreation and/or holiday but that they did not visit during that trip. It was then asked to respondents which of the destinations previously identified they would be most likely to visit. That destination was considered, in the questionnaire, as the main competitor of the Gerês park. Each respondent was required to indicate how several items related to facilities and services to support tourism, tourism attractions and ability to satisfy some motivations, were important in making the Gerês park and its main competitor attractive. People had to answer using a Likert type scale from 1 (not important) to 5 (extremely important). Items included in the questionnaire were selected, as already mentioned, based on the literature review presented in this chapter and in the pilot study conducted already referred. Literature on leisure and tourism presented in previous sections (e.g. Beard and Ragheb, 1983; Crompton, 1979; Lee and Crompton, 1992; Manfredo et al., 1996) was used to identify the items related to the ability to satisfy motivations to be included in the questionnaire. Items regarding tourism attractions and facilities and services to support tourism were also derived from the literature review associated with tourism in general (e.g. Echtner and Ritchie, 1993; Inskeep, 1991) or with protected areas in particular (e.g. Kim, 1998; Ryan and Sterling, 2001). It was also asked to respondents how significant a set of potential constraints were in making the travel to the Gerês park and to the main competitor of the Gerês park, difficult. Respondents had to answer using a Likert type scale from 1 (did not make it difficult) to 5 (made it extremely difficult). Constraints included in the questionnaire were selected, as

Complimentary Contributor Copy

Assessing the Competitiveness of Protected Areas

209

already referred, based on the pilot study conducted and on the literature review on constraints presented in this chapter regarding constraints for participating in specific tourism activities and for visiting protect areas (e.g. Botha et al., 1999; Hudson, 2000; PenningtonGray and Kerstetter, 2002; Tian et al., 1996). Additionally, respondents were requested to indicate if they consulted information about Gerês concerning a set of items related to tourism attractions, facilities and services to support tourism and potential constraints. Finally, respondents had to provide some data regarding their socio-demographic profile and travel behavior during the trip to the Gerês park.

3.1.2. Sampling Procedure Questionnaires were administered to the visitors of Gerês between 15th June 2002 and the end of August 2002. A quota sampling procedure was adopted in order to ensure a similar proportion between the population of visitors of Gerês and the sample, regarding the nationality of the visitors. Questionnaires were administered directly by the researchers in important touristic points of the Gerês National Park. A total of 1115 completed questionnaires was obtained. 3.1.3. Data Analysis Several data obtained are analyzed using univariate data analyses such as frequencies and means. In order to assess the positioning of Gerês park in relation to competitors, in each questionnaire, it was decided to compare, in each question, the perceptions of the respondent regarding the Gerês Park and regarding its main competitor. The perceptions compared related to the attractiveness on Gerês and on its main competitor regarding tourism attractions, facilities and services to support tourism, ability to satisfy motivations and constraints to visit the destination. This is done for all the respondents who had though in, at least, one destination apart from Gerês during the planning of the trip to Gerês. Data is compared using paired-samples t-tests in order to identify statistical significant differences.

3.2. Results and Discussion The majority of respondents are Portuguese (79%). Most of the foreigners are from the following countries: France (29% of the foreigners), Netherlands (18%), Spain (17%), Germany (10%), United Kingdom (8%) and Belgium (7%). The sample is quite balanced regarding the gender (55% of respondents are female), includes a high number of employees (74%) and students (19%) and a big proportion of people between 25 and 44 years old. The majority of the respondents travelled in groups of two to four people (73%), stayed two to seven nights in that park (58%) and more than 40% had already visited the Gerês park before. Only 398 respondents thought in another destination besides the Gerês park when planning the trip to Gerês. This means that only 398 respondents identified a main competitor of the Gerês park. There is a wide diversity among the main competitors identified by respondents. However, the main competitors identified by 43% of the respondents are related to Spain or specific places located in Spain (e.g. Galiza and, specifically, Santiago de Compostela, close to the Gerês park), in Algarve, Alentejo, Azores, Trás-os-Montes, Coast of Alentejo and Serra da Estrela (Figure 2). All the Portuguese areas above referred encompass

Complimentary Contributor Copy

210

Maria João Carneiro and Carlos Costa

Destinations indicated as the main competitor of the Gerês park

protected areas and Serra da Estrela is a protected area itself. Although the majority of these Portuguese areas have a high rural character, the Algarve and the Coast of Alentejo are highly characterized by the presence of beaches. Azores has the particularity of being an archipelago.

Spain Algarve Azores

Alentejo Trás-os-Montes Coast of Alentejo Serra da Estrela 0

10 20 30 40 50 60 70 Number of respondents who indicated the destination as the main competitor of the Gerês park

Figure 2. The main competitors of the Gerês National Park identified by a higher number of respondents.

As far as tourism attractions are concerned, the Gerês park is significantly different from the main competitor identified by each respondent, and is even superior to it regarding natural attractions (such as scenery, fauna and flora) and regarding the lack of congestion and pollution (Table 1, Figure 3). The positioning of Gerês is not so good on cultural attractions where the Gerês has a performance similar to that of its main competitor, or even lower, as far as architecture is concerned. The Gerês park is not very attractive in terms of facilities and supporting services. However, it still has a similar performance than its main competitor regarding restaurants and facilities for providing information and even a higher performance on accommodation and safety. The protected area under analysis is also special because it provides better opportunities for relaxing and for exploring new things than its main competitor. Nevertheless, as far as the ability to satisfy other motivations is concerned, no major differences are detected between Gerês and the competitor. The only exception is that the main competitor offers more opportunities to meet new people. Respondents experience a low level of constraints to visit Gerês, never higher than 1.67 in a scale from 1 (did not make it difficult) to 5 (made it extremely difficult) (Table 2, Figure 4). Visitors to Gerês experience more accessibility and time constraints than financial constraints. However, even the accessibility and time constraints are very low.

Complimentary Contributor Copy

211

Assessing the Competitiveness of Protected Areas

Table 1. Differences between the attractiveness of the Gerês National Park and of its main competitor on a set of facilities and supporting services, tourism attractions and ability to satisfy some motivations Gerês

Main

Gerês vs.

National competitor

Attractions of the destinations scenery

Park

of Gerês

Mean

Mean

main competitor Paired-samples t test t

p

4.65

3.92

13.636

customs

3.28

3.31

-.456

.000 .649

flora and fauna

4.25

3.24

15.899

.000

hospitality

3.39

3.30

1.741

.083

beaches

2.07

3.01

-10.734

.000

historic sites

3.27

3.29

-.246

.806

trails

3.78

2.71

14.861

.000

architecture

3.03

3.27

-3.900

.000

gastronomy

3.37

3.32

.953

.341

rivers and lakes

4.38

3.13

17.006

.000

climate

4.05

4.02

.672

.502

lack of crowds

3.65

3.21

6.887

.000

unpolluted environment

4.61

3.93

11.143

.000

Facilities and supporting services accommodation

3.42

3.23

3.589

.000

safety

3.58

3.40

4.069

.000

facilities for providing information

3.31

3.29

.557

.578

restaurants

3.10

3.09

.362

.717

camping

2.92

2.45

7.369

.000 .000

Destination's ability to satisfy motivations opportunities for resting

4.12

3.74

6.732

opportunities for learning

3.84

3.82

.566

.572

opportunities for meeting new people

3.05

3.14

-2.251

.025

opportunities for contacting local people

3.03

3.10

-1.744

.082

opportunities for avoiding responsabilities

4.36

4.10

6.251

.000

opportunities for experiencing peace and calm

4.31

3.73

9.031

.000

opportunities for experiencing and exploring new things

4.30

4.12

4.053

.000

opportunities for being with friends

3.14

3.12

.511

.609

Note: Likert type scale from 1 "not important" to 5 "extremely important". Cells marked in grey indicate the destination (Gerês park or the main competitor) that is more attractive, in the case of items where there is a statistical significant difference between Gerês and its main competitor (significance level of 5%).

Complimentary Contributor Copy

212

Maria João Carneiro and Carlos Costa

scenery unpolluted environment rivers and lakes opportunities for avoiding responsabilities opportunities for experiencing peace and calm opportunities for experiencing/exploring new things flora and fauna opportunities for resting climate opportunities for learning trails lack of crowds safety accommodation hospitality gastronomy facilities for providing information

customs historic sites opportunities for being with friends restaurants opportunities for meeting new people opportunities for contacting local people architecture camping beaches

1

2

3

4

5

Importance of features in making the destination attractive Gerês National Park

Main competitor of the Gerês park

Figure 3. The attractiveness of the Gerês National Park and of its main competitor on a set of facilities and supporting services, tourism attractions and ability to satisfy some motivations.

The destination features of Gerês about which more respondents collect information are three features related to natural attractions (scenery, rivers and lakes, flora and fauna), two issues associated to accommodation that could be related to potential constraints (type of accommodations available, price of accommodations) an the way to get to the Gerês park, which may also be associated with accessibility constraints (Figure 5). More than 30% of the respondents searched for information about each of these features, what shows the importance of providing information on Gerês regarding these issues.

Complimentary Contributor Copy

213

Assessing the Competitiveness of Protected Areas

Table 2. Differences between the constraints to travel to the Gerês National Park and to its main competitor Gerês

Main

Gerês vs.

National competitor

Constraints to travel to the destination the accommodation at the destination is expensive

Park

of Gerês

Mean

Mean

main competitor Paired-samples t test t

p

1.55

2.09

-9.298

I am too busy

1.59

1.77

-4.055

.000 .000

the transportation infrastructure to the destination is not good

1.59

1.66

-1.310

.191

the travel to the destinations is expensive

1.46

2.10

-10.288

.000

it is difficult to find information on how to get to the destination

1.59

1.48

2.257

.025

the destination is too far from where I live

1.67

2.11

-7.026

.000

I have important things to do

1.29

1.48

-4.877

.000

I do not have enough money

1.35

1.83

-8.788

.000

it is difficult to go to the destination

1.50

1.62

-2.091

.037

it is difficult to find enough time to come to the destination

1.59

1.93

-5.584

.000

Note: Likert type scale from 1 (did not make it difficult) to 5 (made it extremely difficult). Cells marked in grey indicate the destination (Gerês park or the main competitor) that is less attractive, in the case of items where there is a statistical significant difference between Gerês and its main competitor (significance level of 5%). In this case, when the respondents feel more constraints to travel to the main competitor of Gerês than to Gerês, the main competitor is less attractive than Gerês.

the destination is too far from where I live the transportation infrastructure to the destination is not good I am too busy it is difficult to find enough time to come to the destination it is difficult to find information on how to get to the destination the accommodation at the destination is expensive it is difficult to go to the destination the travel to the destinations is expensive I do not have enough money I have important things to do

1.0

1.5

2.0

2.5

Difficulty in traveling to the destination due to specific features Gerês National Park

Main competitor of the Gerês park

Figure 4. The structural constraints to travel to the Gerês National Park and to its main competitor.

Complimentary Contributor Copy

214

Maria João Carneiro and Carlos Costa

scenery rivers and lakes flora and fauna

Features of Gerês about which respondents collected information

the way to get to the destination type of accomodation available

price of the accommodation trails camping areas climate historic sites gastronomy customs price of the travel architecture restaurants

hospitality pollution beaches safety transportation 0

10 20 30 40 50 60 70 Number of respondents who collected information about each feature

Figure 5. Destination features of Gerês about which respondents collected information.

CONCLUSION AND IMPLICATIONS The present chapter differs from previous studies since it proposes a broad methodology for assessing the positioning of destinations, which encompasses a wide range of features related to tourism attractions, facilities and services to support tourism, ability to satisfy motivations and constraints. In this chapter this methodology is also adopted to evaluate the positioning of a protected area – the Gerês park (in Portugal) –, expanding the limited research on the positioning of protected areas. When analyzing the positioning of the Gerês park in relation to competitors, statistical significant differences are detected between Gerês and its main competitor, in all the dimensions used to assess the positioning. Gerês, the

Complimentary Contributor Copy

Assessing the Competitiveness of Protected Areas

215

destination that respondents chose to visit, also reveals to be superior to the competitor in specific issues of all these components. This suggests that the methodology proposed in this chapter, which recommends the assessment of the positioning of destinations based on tourism attractions, facilities and services to support tourism, ability to satisfy motivations and constraints, should be adopted. In the perspective of the respondents, the Gerês park has a more competitive positioning than the competitor. The major strengths of the Gerês park in relation to the competitor are its natural attractions, its lack of pollution and crowds, the opportunities it offers for relaxing and exploring new things and, also, the low constraints felt to travel to Gerês. All these features should be stressed in the promotion of the National Park of Gerês in order to increase its attractiveness. Taking into consideration that data also reveals that the Gerês park has a low performance regarding facilities and supporting services, although this performance is not inferior yet to that of the competitor, those responsible for managing the Gerês park should implement strategies to improve the facilities and supporting services of this protected area. The features of the Gerês park about which more respondents collect information are one of the major strengths of the Park – natural resources – and features associated with potential constraints in visiting the park (related to accommodation and accessibility). These results suggest that the provision of information about the Gerês park has played an important role in enhancing the attractiveness of the Park and in reducing potential constraints for visiting it. In order to increase, even more, the attractiveness of the Gerês park, it is important to continue providing information about the natural attractions of the Park and to also promote the opportunities for relaxation that the Park offers. To ensure that potential visitors will not experience high constraints for visiting the Gerês park, it would be important to provide information about the availability and price of accommodation and especially about the way to get to the destination. This last issue is of particular importance since many respondents searched for this kind of information, but results show that it is much more difficult to find information about the Gerês park than about its main competitor regarding the way to get to the destination. Despite the important insights provided by the study here presented, this study has some limitations. The study was undertaken in a limited time span and would be interesting to carry out a longitudinal study for comparing the positioning of the Gerês Park in different periods of time. It would also be very useful to conduct the study with non-visitors of the Gerês park in order to understand why they do not visit this park and to identify strategies that would enable to attract them to the park. It would also be important to apply this methodology to a wide range of destinations in order to assess the positioning of these destinations in relation to competitors based on a wide range of important factors that determine destinations competitiveness - tourism attractions, facilities and services to support tourism, ability to satisfy motivations and constraints.

REFERENCES Baloglu, S., & McCleary, K. W. (1999). A model of destination image formation. Annals of Tourism Research, 26(4), 868-897.

Complimentary Contributor Copy

216

Maria João Carneiro and Carlos Costa

Beard, J. G., & Ragheb, M.G. (1983). Measuring leisure motivation. Journal of Leisure Research, 15(3), 219-228. Botha, C., Crompton, J. L., & Kim, S.-S. (1999). Developing a revised competitive position for Sun/Lost City, South Africa. Journal of Travel Research, 37(4), 341-352. Cracolici, M. C., & Nijkamp, P. (2008). The attractiveness and competitiveness of tourist destinations: A study of Southern Italian regions. Tourism Management, 30(3), 336-344. Crawford, D.W., & Godbey, G. (1987). Reconceptualizing barriers to family leisure. Leisure Sciences, 9(4), 119-127. Crompton, J. L. (1979). Motivations for pleasure vacation. Annals of Tourism Research, 6(4), 408-424. Deng, J., King, B., & Bauer, T. (2002). Evaluating natural attractions for tourism. Annals of Tourism Research, 29(2), 422–438. Dolnicar, S., Grabler, K., & Mazanec, J.A. (2000). A tale of three cities: perceptual charting for analysing destination images. In A.G. Woodside (Ed.), Consumer Psychology of Tourism, Hospitality and Leisure (Vol 1., pp. 39-62). Wallingford: Cabi Publishing. Dwyer, L., & Kim, C. (2003). Destination competitiveness: determinants and indicators. Current Issues in Tourism, 6(5), 369-414. Echtner, C.M., & Ritchie, J.R.B. (1993). The measurement of destination image: an empirical assessment. Journal of Travel Research, 31(4), 3-13. Haahti, A. J. (1986). Finland’s competitive position as a destination. Annals of Tourism Research, 13(1), 11-35. Holden, A., & Sparrowhawk, J. (2002). Understanding the motivations of ecotourists: the case of trekkers in Annapurna, Nepal. International Journal of Tourism Research, 4(6), 435-446. Hong, S.-K., Kim, J.-H., Jang, H., & Lee, S. (2006). The roles of categorization, affective image and constraints on destination choice: An application of the NMNL model. Tourism Management, 27(5), 750–761. Hsu, T.-K., Tsai, Y.-F., & Wu, H.-H. (2009). The preference analysis for tourist choice of destination: A case study of Taiwan. Tourism Management, 30(2), 288-297. Hudson, S. (2000). The segmentation of potential tourists: constraint differences between men and women. Journal of Travel Research, 38(4), 363-368. ICNF (2012). http://www.icnf.pt/cn/ICNPortal/vPT2007-AP-Geres (assessed in 20 September 2012). IGESPAR (2012). http://www.igespar.pt/pt/patrimonio/pesquisa/geral/ (assessed in 8 September 2012). Inskeep, E. (1991). Tourism Planning: An Integrated and Sustainable Development Approach. New York: Van Nostrand Reinhold. Jackson, E. L. (1993). Recognizing patterns of leisure constraints: results from alternative analyses. Journal of Leisure Research, 25(2), 129-149. Jackson, E. L., & Scott, D. (1999). Constraints to leisure. In E. L. Jackson & T. Burton (Eds.), Leisure Studies: Prospects for the Twenty-First Century (pp. 299-321). State College: Venture Publishing. Jin, X., Bauer, T., & Weber, K. (2010). China’s second-tier cities as exhibition destinations. International Journal of Contemporary Hospitality Management, 22(4), 552-571.

Complimentary Contributor Copy

Assessing the Competitiveness of Protected Areas

217

Kerstetter, D. L., Zinn, H. C., Graefe, A. R., & Chen, P. (2002). Perceived constraints to state park visitation: a comparison of former-users and non-users. Journal of Park and Recreation Administration, 20(1), 61-75. Kim, H.-b. (1998). Perceived attractiveness of Korean destinations. Annals of Tourism Research, 25(2), 340-361. Kim, S. S., Guo, Y., & Agrusa, J. (2005). Preference and positioning analyses of overseas destinations by mainland Chinese outbound pleasure tourists. Journal of Travel Research, 44(2), 212-220. Kim, S.S., Lee, C., & Klenosky, D.B. (2003). The influence of push and pull factors at Korean national parks. Tourism Management, 24(2), 169-180. Kotler, P. (1997). Marketing Management: Analysis, Planning, Implementation and Control (9th ed.), New Jersey: Prentice Hall. Kozak, M., Baloglu, S., & Bahar, O. (2010). Measuring destination competitiveness: Multiple destinations versus multiple nationalities. Journal of Hospitality Marketing and Management, 19(1), 56-71. Lee, T., & Crompton, J. (1992). Measuring novelty seeking in tourism, Annals of Tourism Research, 19(4), 732-751. Manfredo, M.J., Driver, B.L., & Tarrant, M.A. (1996). Measuring leisure motivation: a metaanalysis of the recreation experience preference scales. Journal of Leisure Research, 28(3), 188-213. Marques, C., Reis, E., & Menezes, J. (2010). Profiling the segments of visitors to Portuguese protected areas. Journal of Sustainable Tourism, 18(8), 971–996. Marsinko, A., Zawacki, W.T., & Bowker, J.M. (2002). Use of travel cost models in planning: a case study. Tourism Analysis, 6(3/4), 203-211. McIntosh, R. W., Goeldner, C.R., & Ritchie, J.R.B. (1995). Tourism: Principles, Practices, Philosophies. (7th ed.). New York: John Wiley and Sons. Middleton, V. (1989). Tourist product, In S. Witt, & L. Moutinho (Eds.), Tourism Marketing and Management Handbook (pp.573-576). New York: Prentice-Hall. Middleton, V. T. C., & Clarke, J. (2001). Marketing in travel and tourism. (3rd ed.). Oxford: Butterworth Heinemann. Mill, R. C., & Morrison, A.M. (1998). The Tourism System. (3rd ed.). Dubuque, Iowa: Kendall/Hunt Publishing. Moutinho, L. (1987). Consumer behaviour in tourism. European Journal of Marketing, 21(10), 5-43. Orth, U. R., & Turecková, J. (2002). Positioning the destination product of ‘Southern Moravia’, Journal of Vacation Marketing, 8(3), 247-262. Pennington-Gray, L. A., & Kerstetter, D.L. (2002). Testing a constraints model within the context of nature-based tourism. Journal of Travel Research, 40 (4), 416-423. Raymore, L., Gobey, G., Crawford, D., & von Eye, A. (1993). Nature and process of leisure constraints: an empirical test. Leisure Sciences, 15(2), 99-113. Ritchie, J. R. B., & Crouch, G.I. (2003). The Competitive Destination: A Sustainable Tourism Perspective. Cambridge, MA: CABI Publishing. Ryan, C. (1994). Leisure and tourism – the application of leisure concepts to tourist behaviour – a proposed model. In A. V. Seaton, C. L. Jenkins, R. C. Wood, P. U. C. Duke, M. M. Bennett, L. R. McLellan, & R. Smith (Eds.), Tourism: The State of the Art (pp. 294-307). Chichester: John Willey and Sons.

Complimentary Contributor Copy

218

Maria João Carneiro and Carlos Costa

Ryan, C., & Sterling, L. (2001). Visitors to Litchfield National Park, Australia: a typology based on behaviours. Journal of Sustainable Tourism, 9(1), 61-75. Tian, S., Crompton, J.L., & Witt, P. A. (1996). Integrating constraints and benefits to identify responsive target markets for museum attractions. Journal of Travel Research, 35(2), 3445. Um, S., & Crompton, J.L. (1990). Attitude determinants in tourism destination choice. Journal of Travel Research, 17(3), 432-448. Woodside, A. G., & Lysonski, S. (1989). A general model of traveller destination choice. Journal of Travel Research, 27(4), 8-14.

Complimentary Contributor Copy

In: National Parks Editor: Johnson B. Smith

ISBN: 978-1-62618-934-8 © 2013 Nova Science Publishers, Inc.

Chapter 8

THE DOÑANA NATIONAL PARK (SW SPAIN): FROM THE SEA TO THE FUTURE F. Ruiz1, M. Pozo2, M. Olías1, M. C. Núñez3, M. Abad1, M. I. Carretero4, J. Rodríguez Vidal1, L. M. Cáceres1, J. Prenda5, E. M. Castellanos5, C. J. Luque5, A. Menor5, E. Font6, A. Toscano1 and E. X. García7 1

Departamento de Geodinámica y Paleontología, Universidad de Huelva, Huelva, Spain 2 Departamento de Geología y Geoquímica, Universidad Autónoma de Madrid, Madrid, Spain 3 Departamento de Derecho Público, Universidad de Huelva, Huelva, Spain 4 Departamento de Cristalografía, Mineralogía y Química Agrícola, Universidad de Sevilla, Sevilla, Spain 5 Departamento de Biología Ambiental y Salud Pública, Universidad de Huelva, Huelva, Spain 6 IDL-Faculdade de Ciencias da Universidade de Lisboa, Lisboa, Portugal 7 Departamento de Botánica y Zoología, Universidad de Guadalajara, México

ABSTRACT The Doñana National Park is one of the most important National Parks in Spain. The present-day scenario is the final result of a long geological evolution (~6 Myr), with the ´creation´ of a deep aquifer and numerous geomorphological features that enhance the biodiversity. This socio-ecological system presents numerous threats (invasive species, decrease of aquifer discharges, tourism resorts, sea level rise) and an adequate sustainable development is necessary in order to preserve this special system.

Complimentary Contributor Copy

220

F. Ruiz, M. Pozo, M. Olías et al.

1. INTRODUCTION In the last decade, biodiversity has been one of the most used terms to define the general scenario of life from genes to communities, expressed in a multitude of spatial and temporal scales. According to Noss (1992), biodiversity can be represented as an interlocked hierarchy of elements on several levels of biological organization. Some geographical areas present a very high diversity of species and need a special protection. Biodiversity concerns related to these ecosystems can included (cf. Savard et al., 2000): a) those related to the impact of anthropogenic inputs; b) those related to the increase of biodiversity; and c) those related to the presence of undesirable species. These three ways (and others, if necessary) precise a continuous management and an adequate legal protection. This chapter represents an general overview of the Doñana National Park, one of the most representative areas with high biodiversity in Spain. We start with a general geological/biological overview of the Park. We then review some of the most important threats (biological, geological, human) and finally we propose some actions that can enhance the life diversity in this area.

2. THE DOÑANA NATIONAL PARK 2.1. Birth, Legal Protection and Management of a National Park The Doñana National Park is located on the southwestern Spanish coast, close to the Guadalquivir River mouth (Figure 1, A). The exceptional biodiversity of this zone (see section 2.5) leds to its declaration as National Park in 1969 (Act 2412/1969), with an initial area of 34,600 ha and a later extension to 50,700 ha in 1980. In 1981, this park was included in the Biosphere Reserves and it was declared as Ramsar Site a year later. In the next decade, a first Master Plan for Use and Management (PRUG) was approved (1984) and it receives the European diploma in Management (1985). In 1988, the remarkable ornithological record justified its declaration as a Special Protection Area for birds (ZEPA). In 1989, the periphery of the National Park was declared as Natural Park (54,250 ha) and finally it is a UNESCO World Heritage Site since 1994. Park management is allocated exclusively to the Autonomous Community of Andalusia (Constitutional Court Judgment 194/2004 of November, 10; Royal Decree 712/2006 of June, 9). It is disciplined by the Master Plan for Use and Management approved by decree of the Andalusia (Act 48/2004 of February, 10). The Plan distinguishes four zones: a) Reserve Zone, or areas containing natural values of first order according to their rarity, fragility, biodiversity and scientific interest that require the highest degree of protection; b) Restricted Use Zone, or areas that have a high degree of naturalness and able to withstand a certain level of public use, but may have suffered some degree of human intervention, maintain their natural values in good condition or are in the process of regeneration; c) Moderate Use Zone, or areas dominated by a natural capacity to accommodate more visitors than in previous cases; and d) Special Use Zone, or areas of small size with buildings.

Complimentary Contributor Copy

The Doñana National Park (SW Spain)

Figure 1. Geographical setting and geomorphological features of the Doñana National Park. CM: borehole of Figure 2.

Complimentary Contributor Copy

221

222

F. Ruiz, M. Pozo, M. Olías et al.

2.2. Life Processes of Doñana The life development of the Doñana National Park is greatly controlled by climatology and different hydrodynamic processes. Doñana has a mediterranean climate with oceanic influences, characterized by two seasons (wet: october to mars; dry: april to september), very high variability of annual rainfalls (mean: 550 L/m2) and prolonged periods of drought (Castroviejo, 1993). Mean temperatures vary between 10ºC (december-january) to 25ºC (julyaugust). The main hydrodynamics processes are the fluvial inputs, tidal fluxes, littoral drift currents and the dominant swell. Fluvial imputs of the Guadalquivir River are closely linked to the climatological variations, with an annual average of 164 m3s-1 (Menanteau, 1979) and higher flow rates during the wet season. Tidal regime is mesotidal and semidiurnal, with a mean range of 3.6 m (Borrego et al., 1993). Littoral drif currents follow a clockwise and transport sediments derived mainly from the erosion of Portuguese cliffs and Spanish beaches located to the northwest. These sedimentary contributions and the dominant southwestern swell encourage the creation and expansion of sandy spits (Figure 1, B: Doñana spit), with the development of inner marshlands.

2.3. Skin of a Park: Geomorphological Features The extensive marshes of the Doñana National Park present a flat topography, with some inner depressions occupied by temporary or permanent wetlands so-called ´lucios´ (Figure 1). They are partially isolated from fluvial and/or tidal fluxes by external levees (1-2 m height), although the winter precipitations can cause the flooding of almost all marshlands. The whole system is protected by the Doñana spit, a wide sandy barrier with active dune systems growing toward the southeast. In addition, some “hills” (2-3 m height) are formed by bioclastic or sandy ridges, caused by high-energy events (tsunamis, storms) during the last 5,000 years (Ruiz et al., 2005).

2.4. Growth of a Park: geology and palaeoenvironment The palaeoenvironmental evolution of Doñana National Park has been inferred from the geological record collected in numerous boreholes, with special attention to the PleistoceneHolocene interval (see Ruiz et al., 2010 for a review). Different lithostratigraphic formations have been differentiated in the upper part of this record (Figure 2, A: core CM; modified from Salvany et al., 2011), indicating a growth process over millions of years that continues today. During the Upper Messinian (~6 Myr), a monotonous sequence of greenish to greyish clays (Gibraleón Formation) were deposited in a marine, circalittoral palaeoenvironment. These materials constitute the impermeable base of one of the most important aquifers in southwestern Spain (Almonte-Marismas), connected in some cases with the wetlands mentioned above and therefore very important to the life cycle in this area (see section 3.2). The southwestern Guadalquivir Basin constitutes a wide, shallow bay during the Lower Pliocene (~5-4 Myr), with bottom sediments formed by bioturbated silts and sands. This bottom was occasionally eroded by storms, with the deposit of bioclastic silts (Figure 2, B: Huelva Formation; Civis et al., 1987).

Complimentary Contributor Copy

The Doñana National Park (SW Spain)

223

Figure 2. Synthetic geological record and palaeoenvironmental evolution of the Doñana National Park (modified from Ruiz et al., 2009; Salvany et al., 2011 and Rodríguez Vidal et al., 2011). See Figure 1 for location of borehole CM.

In some cases, these sediments are overlied by the Bonares Formation (bioturbated silts, gravels and conglomerates), indicating a transition to littoral areas. The Lower PleistoceneMiddle to Upper Pleistocene (2.6 Myr-85 kyr) represent a first transition to alluvial

Complimentary Contributor Copy

224

F. Ruiz, M. Pozo, M. Olías et al.

environments (Almonte Formation, Salvany et al., 2011), with yellowish, quartz-rich fine to coarse sands and variable proportions of gravels. These deposits are covered by distal alluvial sediment (Lebrija Formation), with grayish, silty sands and several thick layers of black silty clays with high proportions of organic matter and vegetal remains. During the Late Pleistocene-Late Holocene period (85 kyr-1 kyr), this area was occupied by freshwater/brackish marshes, ponds and aeolian units (e.g. Abalario Formation), with some marine inputs caused by high-energy events (Ruiz et al., 2005; Pozo et al., 2010). This variability of ecosystems was interrupted by the Flandrian Transgression, with a maximum at 7-6.5 kyr (Zazo et al., 1994). At this time, the Doñana National Park and the surrounding areas were flooded (Figure 2, C), with the presence of a lagoon so-called `Lacus Ligustinus´ by romans. This transgression can give us clues to analyze the impact of future sea level rise on the park (see section 3.3. for additional data). The last 6 kyr include an increasing infilling process, with an accelerated growth of spits and new inner marshlands and wetlands (Figure 1). Consequently, the Doñana National Park is the present-day scenario of a geological evolution that started, at least, 6 millions years ago.

2.5. Living in the Park: Biodiversity The biological richness of the Doñana National Park can be explained by its unique geographical location at the crossroads between Europe and Africa and between the Atlantic Ocean and Mediterranean Sea. This is coupled with the presence of three contrasting ecosystems: a) the forest, characterized by a community of Mediterranean scrub; b) dunes, with forests of pines (Pinus pinea), juniper (Juniperus oxycedrus subsp. macrocarpa) and junipers (Juniperus phoenicea subsp. turbinata); and c) wetlands, characterized by periodic flooding.

Figure 3. The imperial eagle (Aquila adalberti).

Complimentary Contributor Copy

The Doñana National Park (SW Spain)

225

This park presents a high biological diversity (>4,000 species), being one of the most important Biosphere Reserve in southern Europe. For example, it includes 2,000 different animals (Figure 3), with more than 1,200 invertebrates (>800 insects) and an important bird diversity (~500 species) (e.g. Fernández et al., 1994; Santos et al., 1998; Máñez and Garrido, 2002). Some invertebrates have been discovered for first time in this area (e.g. Metacyclops planus, Ilyocryptus sordidus, Moina salina). Plants (1,500-1,600 species) are mainly represented by angiosperms (> 1,000 species). Fungi are also very abundant (~400 species). This area is a refuge for numerous vertebrates included on European and even worldscale endangered organisms, such as the Iberian lynx (Lynx pardinus), the imperial eagle (Figure 3: Aquila adalberti), the white headed duck (Oxyura leucocephala) or the marbled teal (Marmorenetta angustirostris).

3. SOME THREATS 3.1. Invasive Species In the last centuries, the Doñana National Park has suffered the introduction of alien invasive species (>100; Martín et al., 2007), causing numerous problems for the global conservation of this sector (e.g. Figure 4). They include aquatic plants (A. filiculoides, C. edulis), trees (Eucalyptus sp.), insects (L. humile), crustaceans (P. clarkii, E. sinensis), fishes (L. gibbosus, M. salmoides, C. carpio) or reptiles (galápago T. scripta). Some of them present a trophic competition with autochthonous species, whereas others hybridize with them.

Figure 4. Spartina densiflora, an invasive species.

3.2. Changes of an Aquifer Many ecosystems of Doñana National Park depend on groundwater from the AlmonteMarismas aquifer, wich has a much larger extension than the Park itself (Figure 5). The water

Complimentary Contributor Copy

226

F. Ruiz, M. Pozo, M. Olías et al.

table is very close to the ground surface, where different plant communities can be distinguished depending on water depth (Muñoz Reinoso and García Novo, 2005; Custodio et al., 2008). The aquifer also feeds a large number of seasonal and permanent ponds and the main streams, such as the Rocina Creek (Figure 5) which forms a rich lush gallery forest. The system’s main source of recharge is direct infiltration from precipitation in the unconfined part of the aquifer. The aquifer resources are between 160 and 210 hm3/year. Discharge occurs in different ways: drainage toward the main streams of the zone, flow to the sea in the coastal zone, evapotranspiration in zones with a shallow water table, and slow flow through the clay layers of the marsh. Finally, by pumpage, which has been relevant since the mid 70’s and is currently estimated around 90 hm3/year (Custodio et al., 2008), mainly for agricultural use. Around 3 hm3/year are withdrawed to supply Matalascañas tourist resort, adjacent to the Park. Pumpings imply the decrease of the rest of aquifer discharges to reach a new equilibrium (Llamas, 1988; Suso and Llamas, 1993). Several works (e.g. Serrano et al., 2008; Gómez Rodríguez, 2009) point out that a number of changes are already occurring in some ecosystems close to the zones where groundwater withdrawals are greater.

Figure 5. Main features of the Almonte-Marismas aquifer (from Olías et al., 2008).

Complimentary Contributor Copy

The Doñana National Park (SW Spain)

227

The main impacts are: 1) decrease of groundwater towards streams and discharge zones, as the Rocina Creek, 2) decrease of hydroperiod and inundation surface of temporal ponds and wetlands, especially near Matalascañas and 3) vegetation changes to more xerophytic communnities in the zones where the height of water table has lowered. These effects can be masked because there is a desconection between the deep aquifer, more transmissive where the groundwater is pumped, and a shallow aquifer which supports the Doñana ecosystems (Figure 5; Olías et al., 2008). Due to the slow hydrodynamic of groundwater the aquifer response to the pumpings is not yet stabilised (Custodio et al., 2008). In agricultural and urban zones groundwater contamination have been detected (Olias et al., 2008). Pollution causes sulphate and nitrate concentration to show stratification in the aquifer, with the highest values in the shallower part. In some points where the pH values are slightly acid (around 6) along with nitrates and sulphates, high concentrations of trace metals from fertilizers are found (Olías et al., 2008). Groundwater quality variations may also have an impact on DNP ecosystems, the natural vegetation in Doñana has adapted to local groundwater chemistry (Manzano and Custodio, 2006), thus any variation in groundwater quality can cause changes in the different ecosystems.

3.3. Sea Level Change and Doñana An important part of this park present a flat topography, with heights around 0-2 m b.s.l. in most cases (Figure 6: Vetalengua).

Figure 6. Topography (in m b.s.l.) of the southern Doñana National Park. See Figure 1 for location.

Complimentary Contributor Copy

228

F. Ruiz, M. Pozo, M. Olías et al.

The inner marshlands are protected from fluvial/tidal inputs by levees not surpassing 1-2 m b.s.l. Consequently, a future sea level rise can provoke the inundation of these areas and the dissapearance of wetlands, very important in the life cycle of this area.

4. FINAL REMARKS: SUSTAINABLE DEVELOPMENT Some general strategies for a sustainable development have been (or may be) proposed: 1. Second Plan for Suitable Development (2007). This Plan includes the protection of key species or a better ecological conectivity between them, an adequate management of solid residues or a decrease of water pollution by anthropogenic inputs (Junta de Andalucía, 2010). 2. Regional Development Plan (RDP; see Oñate et al., 2003 for a review). During 19932000, this Plan injected more than 372 million Euros into the nearest municipalities to Doñana, whereas the new 2000-2006 Structural Funds period of the European Union lacks any specific program for Doñana. This Plan included the promotion and modernization of the agro-food complex or the sustainable use of protected natural spaces. 3. Promotion of regulating services, mainly those related with hydrological regulation. These services present a progressive degradation (Martín et al., 2010). In addition, the influence of new hydrological infrastructures needs an adequate analysis, because some of them may cause a overflooding of the marshes, killing many species that are not adapted to intense winter flooding conditions (Urdiales et al., 2010). 4. Incentives to maintain traditional management practices. The lack of these incentives, due to both market integration and strict conservation policies, has resulted in a territorial matrix in which strict conservation takes place inside the Doñana National Park while land intensification occurs at their borders, establishing a conservation versus development model (Martín et al., 2012). 5. Sustainopreneurship. Business activities may be an example of collaborative interaction between private and public managers in the context of actions for the protection and conservation of the Doñana National Park and the associated natural resources (Gessa and Toledano, 2010).

ACKNOWLEDGMENTS This work was funded by an International Project (FCT PTDC/CTE-GIX/110205/2009), a Spanish DGYCIT Project (CGL2010-15810), an Excellent Project of the Andalusia Board funded by FEDER (SEJ-4770) and five Research Groups of the Andalusia Board (RNM-238, RNM-276, RNM-311, RNM-327 and RNM-349).

Complimentary Contributor Copy

The Doñana National Park (SW Spain)

229

REFERENCES Borrego, J., Morales, J. A., Pendón, J. G. (1993). Holocene filling of a estuarine lagoon along the mesotidal coast of Huelva: the Piedras River mouth, Southwestern Spain. J. Coast. Res., 9, 242-254. Castroviejo, J. (1993) Memoria. Mapa del Parque Nacional de Doñana. Consejo Superior de Investigaciones Científicas (CSIC) y Agencia de Medio Ambiente (AMA) de la Junta de Andalucía. Madrid. Civis, J., Sierro, F. J., González Delgado, J. A., Flores, J. A., Valle, M. F. (1987). El Neógeno marino de la provincia de Huelva: Antecedentes y definición de las unidades litoestratigráficas. In: Paleontología del Neógeno de Huelva (W Cuenca del Guadalquivir) (Civis, J., ed.), 9-23. Universidad de Salamanca, Spain. Custodio, E., Manzano, M., Montes, C. (2008). Perspectiva general del papel y gestión de las aguas subterráneas en el área de Doñana, Sudoeste de España. Boletín Geológico y Minero, 119, 81-92. Fernández, C., Herrera, M., Sánchez, F. J., Ariza, J.C. (1994). Inventario de las especies de peces del Parque Nacional de Doñana. Biología, ecología y conservación. Memoria inédita. ICONA. Gessa, A., Toledano, N. (2010). Turismo, emprendimiento y sostenibilidad en los espacios naturales protegidos. El caso de Andalucía, España. Estudios y Perspectivas en Turismo, 20, 1154-1174. Gómez Rodríguez, C. (2009). Condicionantes ecológicos de la distribución de anfibios en el Parque Nacional de Doñana. Ph.D. Thesis. University of Salamanca. Junta de Andalucía (2010). Memoria Ambiental II Plan de Desarrollo Sostenible de Doñana. Consejería de Medio Ambiente. 21 pp. Llamas, M. R. (1988). Conflicts between wetland conservation and groundwater exploitation: two case histories in Spain. Environ. Geol. Water Sci., 11, 241-251. Manzano, M., Custodio, E. (2006). The Doñana aquifer and its relation with the natural environment. In: Garcia-Novo F, Marín-Cabrera C (eds), Doñana: Water and Biosphere, Madrid, Spain, 141-150. Máñez, M., Garrido, H. (2002). Avifauna. In: Parque Nacional de Doñana (García, V., coord.), 231-246. Canseco Editores. Martín, B., García, P., Alcorlo, P., Montes, C. (2007). El valor económico como indicador de la amenaza de las especies invasoras. El caso de los Parques Nacional y Natural de Doñana. In: Invasiones biológicas: un factor del cambio global (GEIB, eds.), 40-75. GEIB, serie técnica nº 3. Martín, B., García, M., Gómez, E., Montes, C. (2010). Evaluación de los servicios de los ecosistemas del sistema socio-ecológico de Doñana. Rev. Cátedra UNESCO sobre Desarrollo Sostenible de la UPV/EHU, 4, 91-112. Martín, B., Iniesta, I., García, M., Palomo, I., Casado, I., García, D., Gómez, E., Oteros, E., Palacios, I., Willaarts, B. González, J. A., Santos, F., Onaindia, M., López, C., Montes, C. (2012). Uncovering Ecosystem Service Bundles through Social Preferences. PloS ONE, 7, e38970.

Complimentary Contributor Copy

230

F. Ruiz, M. Pozo, M. Olías et al.

Menanteau, L. (1979). Les Marismas du Guadalquivir. Exemple de transformation d’un paysage alluvial au cours du Quaternaire récent. These 3er Cycle, Paris-Sorbonne University, Paris, 154 pp. Muñoz Reinoso, J. C., García Novo, F, (2005), Multiscale control of vegetation patterns: the case of Doñana (SW Spain). Landscape Ecology, 20, 51-61. Noss, R. (1992). The Wildlands Project land conservation strategy. Wild Earth, 1, 10-25. Olías, M., González, F., Cerón, J. C., Bolívar, J. P., González-Labajo, J., García-López, S. (2008). Water quality and distribution of trace elements in the Doñana aquifer. Environ. Geol., 55, 1555-1568. Oñate, J. J., Pereira, D., Suárez, F. (2003). Strategic Environmental Assessment of the Effects of European Union’s Regional Development Plans in Doñana National Park (Spain). Environ. Management, 31, 642-655. Pozo, M., Ruiz, F., Carretero, M. I., Rodríguez Vidal, J., Cáceres, L. M., Abad, M., GonzálezRegalado, M. L. (2010). Mineralogical assemblages, geochemistry and fossil associations of Pleistocene-Holocene complex siliciclastic deposits from the Southwestern Donana National Park (SW Spain): A palaeoenvironmental approach. Sed. Geol., 225, 1-18. Rodríguez Vidal, J., Ruiz, F., Cáceres, L. M., Abad, M., González-Regalado, M. L., Pozo, M., Carretero, M. I., Monge, A. M., Gómez, F. (2011). Geomarkers of the 218-209 BC Atlantic tsunami in the Roman Lacus Ligustinus (SW Spain): A palaeogeographical approach. Quat. Int., 242, 201-212. Ruiz, F., Rodríguez-Ramírez, A., Cáceres, L. M., Rodríguez Vidal, J., Carretero, M. I., Abad, M., Olías, M., Pozo, M. (2005). Evidence of high-energy events in the geological record: Mid-Holocene evolution of the southwestern Doñana National Park (SW Spain). Palaeogeog., Palaeoclimatol., Palaeoecol., 229, 212-229. Ruiz, F., González-Regalado, M. L., Abad, M., Civis, J., González Delgado, J. A., Muñoz, J. M., Pendón, J. G., Toscano, A. (2009). Impact of storms on Pliocene benthic foraminiferal assemblages of southwestern Spain. Ameghiniana, 46, 345-360. Ruiz, F., Pozo, M., Carretero, M. I., Abad de los Santos, M., González-Regalado, M. L., Muñoz Pichardo, Rodríguez Vidal, J., Cáceres, Puro, L. M., Pendón, J. G., Prudêncio, M. I. and Dias, M. I. (2010). Birth, Evolution and Death of a Lagoon: Late Pleistocene to Holocene Palaeoenvironmental Reconstruction of the Doñana National Park (SW Spain). In: Lagoons: Biology, Management and Environmental Impact (Friedmann, A. G., ed.), 371-396. Nova Science Publishers, New York. Salvany, J. M., Cruz, J., Mediavilla, C., Rebollo, A. (2011). Chronology and tectonosedimentary evolution of the Upper Pliocene to Quaternary deposits of the lower Guadalquivir foreland basin, SW Spain. Sed. Geol., 241, 22-39. Santos, S., Carretero, M. A., Llorente, G. A., Montori, A. (1998). Inventario de las áreas importantes para anfibios y reptiles de España. Colección técnica. Publicaciones del Organismo Autónomo de Parques Nacionales. Ministerio de Medio Ambiente. Savard, J. P. L., Clergeau, P., Mennechez, G. (2000). Biodiversity concepts and urban ecosystems. Landscape and Urban Planning, 48, 131-142. Serrano, L., Esquivias-Segura, M. P., Zunzunegui, M. (2008). Long term hydrological changes over a seventeen-year period in temporary ponds of the Doñana N.P. (SW Spain). Limnetica, 27, 65-78. Suso, J., Llamas, M. R. (1993). Influence of groundwater development on de Doñana National Park ecosystems (Spain). J. Hydrol., 141, 239-270.

Complimentary Contributor Copy

The Doñana National Park (SW Spain)

231

Urdiales, C., García, D., Valero, A., Fernánez, J. M. (2010). Seguimiento de la inundación en la marisma de Doñana: resultados del ciclo 2009/2010 y efecto del dique de la Montaña del Río en el proceso de inundación. In: Tecnologías de la Información Geográfica: La Información Geográfica al servicio de los ciudadanos. Secretariado de Publicaciones (Ojeda, J., Pita, M. F., Vallejo, I., eds.), 1146-1156. Universidad de Sevilla. Zazo, C., Goy, J. L., Somoza, L., Dabrio, C. J., Belloumini, G., Improta, S., Lario, J., Bardají, T., Sillva, P. G. (1994). Holocene sequence of sea-level fluctuations in relation to climatic trenes in the Atlantic-Mediterranean linkage coast. J. Coastal Res., 10, 935-944.

Complimentary Contributor Copy

Complimentary Contributor Copy

INDEX A access, viii, 8, 9, 10, 11, 25, 26, 27, 28, 29, 31, 35, 47, 48, 53, 54, 55, 56, 58, 64, 66, 68, 69, 70, 87, 89, 106, 110, 120, 122, 124, 157, 158, 172, 185, 187, 192, 193, 194, 198, 204 accessibility, 14, 203, 204, 205, 206, 207, 210, 212, 215 accommodation(s), 203, 204, 205, 206, 207, 210, 212, 215 accountability, 14, 31, 48, 58, 89, 95, 109, 121, 123, 124 acid, 227 adaptability, 108, 109, 110, 124 adaptation, 111 adjustment, 13, 167, 170 administrators, 8, 58 advocacy, 23 aesthetic(s), 193, 206 Africa, ix, 2, 5, 8, 14, 18, 20, 23, 28, 31, 36, 37, 40, 41, 42, 43, 44, 50, 53, 54, 56, 66, 68, 73, 102, 103, 126, 127, 155, 156, 158, 174, 175, 176, 177, 224 age, 113, 116, 180 agencies, 6, 17, 26, 34, 65, 91, 123, 157, 180 agriculture, 5, 6, 7, 8, 11, 12, 13, 16, 20, 26, 29, 32, 33, 34, 37, 38, 63, 66, 68, 89, 136, 156, 173 AIDS, 2, 15 Alaska, 175 algorithm, 131 alienation, 6, 8, 22, 66 ambivalence, 198 American culture, 183, 193 amphibians, 14 ancestors, 187 animal husbandry, 98 annuals, 156 anxiety, 204

appraisals, 97 aquifers, 222 arbitration, 43 arrest(s), vii, 1, 86, 87 artifacts, vii, 1, 30 Asia, 28, 43 assessment, 39, 89, 90, 107, 108, 109, 110, 130, 150, 151, 153, 154, 160, 202, 204, 215, 216 assets, 31, 35, 172, 194 asymmetry, 34, 146 attachment, 20 attitudes, 20, 35, 39, 42, 49, 50, 73, 84, 85, 86, 87, 89, 90, 100, 101, 119, 151, 198 authority(s), 6, 9, 10, 11, 12, 20, 22, 23, 24, 25, 26, 34, 48, 49, 56, 59, 60, 61, 69, 71, 74, 80, 84, 85, 86, 87, 89, 100, 106, 114, 115, 119, 122, 124, 188, 189, 192 awareness, 16, 29, 95, 96

B banks, 185 barriers, 136, 204, 207, 216 barter, 21 base, viii, 23, 27, 47, 98, 160, 181, 188, 189, 222 basic needs, 49 Belgium, 209 belief systems, 31 beneficiaries, 22, 33, 98 benefits, 4, 19, 21, 22, 24, 28, 34, 35, 49, 50, 58, 64, 65, 66, 69, 79, 85, 87, 89, 94, 98, 99, 107, 110, 115, 117, 118, 123, 156, 157, 158, 170, 171, 172, 173, 176, 218 Bhagwati, 31, 37 bias, 26, 75, 94, 95, 123 biogas, 118, 123 biomass, 14, 167 biosphere, 3

Complimentary Contributor Copy

Index

234 biotechnology, 36 birds, 10, 14, 55, 154, 208, 220 blame, 60 Bolivia, 33, 37 bone, 7 boreal forest, 153 boreholes, 222 Botswana, 160, 174 bounds, 21, 24, 181 Brazil, 62 breakdown, 58 breeding, 114 brothers, 191 browser, 168, 171 browsing, 167, 168 budding, 183, 194 buffalo, 167, 171 Bureau of Land Management, 179, 189 burn, 17, 20, 30 businesses, 28, 194

C caliber, 20 Cameroon, v, vii, 1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 16, 17, 18, 20, 22, 23, 24, 26, 27, 28, 29, 30, 31, 32, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45 CAMPFIRE, 158 candidates, 132, 138 capacity building, 160 carbon, 28, 35 case study(s), vii, ix, 1, 4, 14, 41, 101, 107, 109, 112, 124, 132, 136, 137, 138, 142, 144, 147, 148, 150, 151, 152, 153, 155, 157, 158, 159, 160, 166, 170, 173, 199, 216, 217 cash, 7, 51, 68, 89, 106, 119 cash crops, 7, 68 catastrophes, 62, 73 catastrophic failure, 17 categorization, 216 cattle, 25, 68, 86, 119, 120, 157, 166, 184, 188 CBD, 47, 62, 102 CBNRM, 2, 10, 106, 158 Central African Republic, 2, 17, 29 Central Europe, 1 certification, 29 Chad, 27, 28, 30 challenges, 6, 9, 49, 54, 59, 73, 95, 97, 100, 108, 153, 171, 172, 181, 193, 194 Chicago, 37 children, 23, 24, 85 China, 28, 216

CIA, 4 city(s), 183, 187, 192, 198, 202, 205, 216 citizens, 11, 48, 110, 206 civil law, 8 civil service, 13 civil society, 66, 97 civil war, 57 civilization, 196 clarity, 123, 172 classification, 3, 4, 133, 149, 204, 206 climate, 10, 23, 28, 35, 41, 111, 130, 152, 204, 222 climate change, 23, 28, 35, 152 clusters, 33, 37 coal, 191 cobalt, 30 cocoa, 7, 18, 20, 30 codes, 107 codes of conduct, 107 coffee, 18, 30, 56 cognition, 116 coherence, 48 Cold War, 191 collaboration, 20, 34, 65, 71, 89, 94, 110, 185 colleges, 16 collusion, 20 colonisation, 40 colonization, 9, 13 commercial, 6, 7, 11, 13, 18, 20, 29, 30, 32, 34, 35, 56, 156, 157, 172, 173 commercial hunting, 11, 18, 20, 29 commodity, 193, 194 communication, 109, 111, 120 community conflict, 101 community relations, 70, 74 community support, 89, 119 compensation, 11, 28, 29, 48, 52, 56, 76, 80, 82, 89, 106, 110, 114, 115, 119 competing interests, 196 competition, 27, 188, 225 competitive advantage, x, 202 competitiveness, x, 201, 202, 203, 204, 205, 206, 207, 215, 216, 217 competitors, x, 201, 202, 207, 209, 210, 214, 215 complementarity, 130, 149, 152 complexity, 59, 85, 131, 136 compliance, 106, 107, 108 composition, 113, 159 concordance, ix, 130 conflict, viii, 4, 6, 8, 10, 20, 21, 22, 23, 26, 27, 36, 42, 47, 48, 49, 51, 53, 54, 57, 60, 61, 69, 73, 74, 75, 79, 84, 85, 86, 87, 88, 105, 106, 107, 109, 110, 111, 113, 118, 122, 124, 180, 190, 191, 192 conflict resolution, 36, 53, 54, 109, 124

Complimentary Contributor Copy

Index Congo, 5, 17, 18, 29, 42, 60 congress, 62, 151, 185 congruence, 131, 143, 145, 146 connectivity, 109, 206 consciousness, 63 consensus, 62, 130, 207 conservation initiatives, viii, 9, 10, 105, 106, 108, 109, 110, 111, 112, 124 conservation programs, 14 conservation status, viii, 47, 151, 154 conserving, vii, ix, 2, 22, 34, 64, 106, 154, 155, 158, 160, 164 Constitution, 11, 64 construction, 16, 22, 28, 69, 70 consulting, 110 consumers, 28, 170, 202 consumption, 12, 68, 113 contamination, 227 contingency, 78 controversial, 192 controversies, x, 179 convention, 35, 102 Convention on Biological Diversity, 102 convergence, 66, 99 conversations, 74, 75 conviction, 194 cooking, 33 cooperation, viii, 16, 18, 105, 124, 158, 190, 192 coordination, 24, 30, 109, 111, 115, 122, 124, 125 copper, 191 correlation, 78, 131, 145, 146, 150 correlation coefficient, 131 correlational analysis, 77 corruption, 13, 14, 20, 35, 98 cost, 34, 50, 68, 86, 117, 193, 207, 217 cotton, 183 covering, 3, 108, 109, 111, 161, 188 craving, 194 criminality, 54 crises, 26 criticism, 49 crop(s), 10, 12, 20, 26, 33, 53, 56, 57, 59, 60, 61, 68, 69, 75, 76, 79, 80, 82, 84, 85, 86, 87, 88, 100, 119 crowds, 215 crown, 56, 58 cultivation, ix, 7, 17, 20, 23, 28, 30, 42, 155, 156, 157 cultural heritage, 208 cultural norms, 9 cultural tradition, 49, 68 culture, x, 8, 11, 21, 34, 40, 54, 172, 179, 183, 187, 188, 194, 196, 206 cure, 15

235

customers, 28

D data analysis, 117 data collection, 120 data set, 131, 135, 136, 137, 139, 141, 142, 143, 144 database, 163 decentralisation, 109 decentralization, 23, 34, 35 decision makers, 149 decision-making process, 22, 52, 86, 90, 91, 98, 111, 206 deep aquifer, x, 219, 227 deforestation, 5, 7, 10, 35, 39 degradation, ix, 5, 7, 10, 13, 22, 25, 26, 40, 42, 63, 155, 156, 160, 170, 188, 228 delegates, 26 Democratic Republic of Congo, 5, 66 democratization, 10 Department of Agriculture, 166, 175 dependent variable, 78 deposits, 30, 183, 189, 191, 224, 230 depression, 188 deprivation, 8 depth, 6, 116, 226 destruction, 20, 23, 32, 36, 50, 63, 156 developed countries, 39, 62 developing countries, 21, 49, 62, 106 development policy, 66, 99 devolution, 13, 31, 117, 121 diet, 121, 167, 168 disaster, vii, 2, 20 discharges, x, 219, 226 displacement, 6, 26, 27, 172 dissatisfaction, 118 distribution, ix, 21, 50, 64, 66, 78, 99, 110, 116, 117, 118, 123, 124, 130, 131, 132, 136, 149, 150, 152, 159, 163, 164, 165, 173, 230 diversification, 9 diversity, vii, viii, 1, 9, 11, 14, 17, 19, 25, 34, 37, 44, 47, 62, 102, 106, 107, 132, 136, 151, 152, 153, 154, 164, 167, 180, 207, 209, 220, 225 DOI, 187, 188 domestication, 35 dominance, 187 donors, 13, 24, 27, 89 Douala, 25 draft, 40, 72 drainage, 191, 226 drawing, 35 dream, 24 drinking water, 168

Complimentary Contributor Copy

Index

236 drought, 26, 27, 28, 180, 187, 188, 198, 222

E Earth Summit, 34 ecological requirements, 165 ecological systems, 111 ecology, 37, 102 economic activity, 22 economic consequences, 106 economic crisis, 13 economic development, ix, 13, 24, 40, 48, 63, 99, 106, 130, 131, 132, 136, 150, 160, 192 economic downturn, 13 economic growth, 30 economic incentives, 103 economic indicator, 87 economic relations, 31 economic well-being, 49 economics, 40, 42, 49, 101, 151, 180, 198 ecosystem(s), viii, 2, 9, 13, 14, 17, 18, 24, 27, 33, 38, 47, 130, 153, 159, 170, 193, 220, 224, 225, 227, 230 eco-tourism, 89 Ecotourism, 41, 103, 175, 176 education, 4, 16, 34, 35, 87, 88, 90, 114 Egypt, 204 elaboration, 120 election, 193 electricity, 22, 27, 28, 123, 192 elephants, 10, 17, 21, 29, 30, 56, 60 emergency, 26 emigration, viii, 2, 33 employees, 118, 209 employment, 69, 114, 157 empowerment, 109, 110, 111, 124 endangered, viii, 13, 14, 16, 17, 24, 28, 47, 68, 111, 225 endangered species, 17 endowments, 108 energy, 5, 28, 63, 88, 222, 224, 230 energy consumption, 5 enforcement, 29, 33, 60, 84, 85, 87, 89, 107 entrepreneurs, 191 environment(s), 4, 11, 17, 22, 24, 34, 35, 36, 37, 44, 62, 63, 64, 66, 101, 136, 156, 158, 161, 180, 193, 194, 198, 204, 206, 207, 224, 229 environmental change, 152 environmental degradation, 4, 22, 23, 61, 63, 158 environmental impact, vii environmental issues, 11 environmental management, 23, 109 environmental movement, 62

environmental policy, 131, 180, 199 environmental protection, 10, 21, 63 environmental quality, 64 environmental regulations, 31 EPS, 197 equality, 124 equilibrium, 226 equipment, 55, 88, 204, 207 equity, 13, 32, 42, 48, 109, 110, 111, 124 erosion, 63, 119, 130, 156, 160, 222 ethnic groups, 32, 112 EU, 2, 16, 18, 29, 44 Europe, 151, 154, 191, 224, 225 European Commission, 18 European Union, 2, 18, 44, 228, 230 evapotranspiration, 226 evidence, 6, 16, 51, 60, 63, 66, 72, 75, 94, 122, 158 evolution, vii, x, 1, 6, 9, 153, 219, 222, 223, 224, 230 exclusion, 10, 24, 25, 57, 157, 169, 172, 173 exercise, 5, 19, 31 exhaustible natural resources, 63 expertise, 13, 24, 26, 49, 194 exploitation, 3, 6, 7, 8, 9, 13, 14, 18, 19, 25, 27, 29, 32, 34, 35, 50, 89, 157, 193, 229 extinction, 3, 130, 152, 153 extraction, viii, 11, 14, 63, 70, 102, 105, 122, 176

F facilitators, 116 fairness, 48, 109 faith, 180, 185 families, 8, 185, 188 family members, 31 farmers, 10, 17, 28, 56, 57, 60, 63 farmland, 69 farms, 27, 156, 157, 185, 188 fauna, 14, 24, 25, 26, 29, 185, 203, 206, 207, 208, 210 FDR, 189 fears, 193, 194 federal authorities, 180 federal government, 192 feelings, 193 fertilizers, 227 financial, viii, 24, 28, 62, 68, 69, 105, 109, 110, 117, 204, 205, 207, 210 financial crisis, 28 financial resources, 24, 62, 110, 117, 204 Finland, 204, 216 firearms, 21 fires, 69, 76, 154 fish, 14, 18, 55, 70, 88

Complimentary Contributor Copy

Index fisheries, 11 fishing, 25, 55, 85 flaws, 74 flexibility, 109, 111, 122, 124, 130 flooding, 222, 224, 228 floods, 73, 180 flora, 14, 25, 26, 130, 136, 151, 152, 185, 203, 206, 207, 208, 210, 212 flora and fauna, 212 flowers, 168 fluctuations, 231 focus groups, 116 folklore, 55 food, 16, 26, 34, 49, 51, 59, 68, 70, 120, 159, 168, 203, 206, 207, 228 food products, 26 forbs, 168 force, 33 Ford, 189 forest fire, 75, 136 forest management, 5, 11, 34, 37, 38, 41, 42, 56, 57, 58, 62, 73, 109 forest resources, 3, 8, 11, 14, 17, 18, 20, 21, 29, 30, 56, 64, 68, 69, 71, 85, 89, 101, 121 formation, 51, 121, 132, 137, 171, 215 formula, 161 fragility, 220 France, 209 freshwater, 150, 224 friendship, 207 fruits, 16, 17, 50, 120, 121, 168 funding, 24, 49, 88, 89, 97, 98, 99, 120, 123 funds, 33, 45, 50, 86, 89, 94, 95, 97, 114, 117, 122, 176

G Gabon, 17, 39 game farming, vii, ix, 155, 156, 158, 160, 170, 171, 172, 173 GEF, 88, 89 genes, 220 Geographic Information System, 2, 41, 176 geography, 180 geology, 185, 222 Germany, 204, 209 GIS, 2, 4, 161, 162, 163, 176 Global Environment Facility, 37, 160 global scale, 149 globalization, 10, 193 governance, vii, 1, 9, 23, 24, 31, 42, 48, 49, 51, 52, 54, 58, 62, 63, 64, 65, 86, 89, 90, 91, 94, 95, 96, 97, 98, 99, 100, 109

237

governments, 9, 20, 28, 48, 58, 158 grades, 132, 140, 141 grants, 11, 94, 106, 171 graph, ix, 130, 132, 133, 135, 137, 138, 142, 143, 144, 145, 146, 147, 148, 149, 150 grass(s), 49, 114, 119, 120, 121, 166, 168, 184, 188 grasslands, 111, 120, 168 grassroots, 90, 98 gravity, 88, 91 grazers, 166, 167, 168, 170 grazing, 22, 26, 27, 57, 121, 159, 167, 180, 184, 187, 188, 189, 191, 194 Great Britain, 39, 175 Great Depression, 188 Greece, 129, 151, 153, 154, 204 groundwater, 225, 227, 229, 230 growth, 10, 12, 23, 29, 45, 50, 60, 187, 222, 224 guidance, 196 guidelines, 65, 113, 117, 121, 124 guiding principles, 65 guilty, 60 Guinea, 14, 17

H habitat(s), ix, 11, 12, 13, 17, 29, 69, 114, 155, 156, 159, 160, 162, 164, 165, 166, 167, 168, 169, 171, 172, 173 Haiti, 49 happiness, 203 hardwood forest, 28 harmony, 22 harvesting, 33, 49, 50, 54, 58, 59, 71, 78, 89, 103, 117, 121, 156 health, 15, 30, 34, 49, 64, 69, 88, 114, 159, 193, 207 health care, 34, 114 height, 222, 227 hemisphere, 163 herbal medicine, 69 herpetofauna, 14, 152 heterogeneity, 27, 152 highlands, 68, 101 highways, 193, 197 history, vii, x, 9, 60, 76, 179, 180, 181, 192, 194, 198, 206 HIV, 15 Holocene, 222, 224, 229, 230, 231 homes, 158 homogeneity, 24 host, 187, 193 hostility, 20, 23, 85, 192 hotels, 204 hotspots, 41, 131, 152, 153, 154

Complimentary Contributor Copy

Index

238

household income, 88, 95 housing, 90 human, vii, viii, 1, 3, 4, 7, 12, 13, 16, 17, 20, 22, 23, 26, 27, 31, 42, 47, 48, 60, 64, 69, 99, 109, 111, 119, 136, 156, 159, 162, 180, 184, 194, 220 human activity, 13, 69 Human Development Report, 44 human right(s), 4, 23, 31, 48, 99 human security, 22 human welfare, 22 Hungary, 1 Hunter, 199 hunter-gatherers, 32 hunting, 3, 11, 12, 13, 16, 17, 18, 21, 25, 26, 28, 29, 30, 55, 57, 59, 68, 74, 170, 171, 172, 173 hypothesis, 74

I ICDPs, 2, 10, 13, 44, 49, 50 ID, 40, 199 ideal(s), 20, 152, 194 identification, 65, 95, 97, 108, 131, 153, 164, 206 identity, 26, 71, 172, 187, 193 ideology, 10, 22, 27 illusion, 189 image(s), 40, 161, 162, 163, 165, 203, 206, 215, 216 imagery, 41, 159, 161 immigration, 9, 34 impact assessment, 90 Impact Assessment, 27, 128 improvements, 73, 188 incidence, 26, 168 income, 11, 29, 33, 34, 48, 50, 53, 59, 64, 68, 69, 72, 88, 89, 106, 107, 110, 118, 156, 171, 172, 173 increased workload, 24 independence, 6, 10, 54, 58, 61, 62, 168, 188 India, 28, 34 indigenous knowledge, 20, 22, 44 indigenous peoples, 12, 22, 23, 43 individual perception, 59 individualism, 188 individuals, viii, 8, 20, 21, 31, 52, 55, 59, 61, 74, 76, 80, 81, 85, 86, 87, 94, 105, 107, 115, 169, 204 Indonesia, 128 industry, 28, 188, 189, 191 information exchange, 111 information sharing, 111 infrastructure, 13, 114, 172, 192, 204 ingredients, 159 injury, 69, 119, 188, 193 insects, 225 insecurity, 26, 32

institutional reforms, 10 institutions, viii, 9, 22, 23, 24, 26, 31, 34, 51, 57, 58, 65, 66, 99, 105, 106, 107, 108, 109, 111, 115, 116, 117, 120, 121, 124, 158, 173, 194, 196 insurgency, 58 Integrated Conservation-Development Projects, 126 integration, viii, 4, 109, 129, 130, 132, 142, 149, 150, 193, 228 integrity, 9, 16, 29, 59, 63 interdependence, 133 interference, 192 international trade, 10 interrelations, 130 intervention, 35, 61, 86, 220 invertebrates, 225 investment(s), 10, 13, 22, 23, 27, 66, 156 Iowa, 217 irrigation, 114, 185, 194 islands, ix, 36, 129, 132, 136, 144, 145, 146, 147, 148, 150, 157 isolation, 183, 187, 188, 189, 191, 193, 194, 196, 197 issues, viii, 19, 35, 47, 48, 50, 52, 54, 60, 62, 76, 82, 84, 85, 90, 94, 97, 98, 99, 100, 111, 115, 116, 117, 118, 120, 181, 198, 212, 215 Italy, 151

J Japan, 204 jaundice, 36 joint ventures, 158 jurisdiction, 193 justification, 23, 30

K Kenya, 7, 22, 42, 44, 118, 125, 175, 176 kill, 10, 17, 21, 34, 60, 121 kinship, 33 Korea, 206

L lakes, 212 land acquisition, 32 land disposal, 185 land tenure, vii, ix, 2, 6, 7, 9, 23, 26, 31, 32, 33, 42, 155 Land Use Policy, 151

Complimentary Contributor Copy

Index landscape(s), 22, 27, 36, 39, 57, 65, 131, 149, 152, 157, 158, 180, 181, 185, 188, 190, 193, 194, 196, 204 Late Pleistocene, 224, 230 law enforcement, 29, 54, 60, 65, 71, 74, 75, 76, 79, 80, 81, 82, 83, 84, 85, 86, 87, 100, 113 laws, 9, 11, 14, 21, 22, 24, 29, 31, 32, 54, 57, 58, 66, 99, 107 laws and regulations, 57 lead, 12, 29, 30, 86, 87, 110, 156, 189, 192 leadership, 13, 31, 63, 194 learning, 53, 100, 101, 109, 110, 111, 121, 124 learning process, 124 legal protection, 220 legend, 192 legislation, 4, 11, 13, 21, 24, 30, 65, 113 leisure, 203, 204, 208, 216, 217 levees, 222, 228 Liberia, 42 life cycle, 222, 228 lifetime, 21 light, 10, 160, 173, 206 Likert scale, 75, 76 Lion, 36 litigation, 7 livestock, ix, 10, 12, 27, 33, 37, 60, 65, 88, 113, 114, 119, 155, 156, 157, 159, 166, 183, 184, 187, 188 loans, 98 lobbying, 185, 191 local community, 20, 22, 31, 59, 64, 65, 74, 75, 76, 78, 88, 89, 101, 171, 172 local government, 31, 58, 90, 93, 94, 95, 97 logging, 5, 7, 10, 13, 21, 23, 25, 29, 30, 32, 33, 34, 35, 191 logistics, 25 longitudinal study, 215 love, 194

M machinery, 23 major decisions, 51, 121 majority, 26, 28, 35, 91, 118, 202, 203, 208, 209, 210 malaria, 26 Malaysia, 37 mammal(s), 14, 17, 20, 35, 159, 165, 166 man, 33, 77, 86, 136 management committee, 26 manganese, 30 mania, 157 manpower, 24 mapping, 151, 152, 164, 173, 185 marginalization, 6, 32

239

marketing, 17, 89, 193 marriage, 187 marsh, 226 Maryland, 102 masculinity, 21 mass, 193 materials, 13, 16, 49, 68, 69, 98, 222 matrix, 137, 148, 228 matter, 66, 193 Mauritius, 88 measurement(s), 41, 108, 216 meat, 29, 30, 57, 70, 89, 188 Mediterranean, ix, 129, 130, 132, 136, 145, 147, 151, 152, 224, 231 membership, 33 messages, 94 meta-analysis, 217 metals, 227 methodology, x, 201, 202, 205, 214, 215 metropolitan areas, 196 Mexico, 44, 126 Middle East, 28 migrants, 25, 113 migration, 29, 33, 120, 187 military, 63 mission(s), 182, 183, 184, 189, 192, 193 Missouri, 152 misunderstanding, 60 misuse, 26, 55, 59, 94 models, 10, 19, 203, 204, 217 modernization, 228 moisture, 163 monopoly, 35, 158 moratorium, 27 mortality, 45 mosaic, 18 Moscow, 199 motivation, 216, 217 Mozambique, 32, 158 multinational companies, 10 multiple factors, 159 museums, 203 mutual respect, 34 mutuality, 65

N Namibia, 49, 158, 174, 175, 176, 177 National Academy of Sciences, 128 national interests, 181 National Park Service, x, 40, 179, 191, 198

Complimentary Contributor Copy

Index

240

national parks, vii, viii, 1, 2, 3, 4, 25, 34, 47, 48, 49, 50, 51, 53, 54, 57, 59, 61, 88, 89, 90, 98, 99, 100, 101, 113, 205, 206, 217 national policy, 53, 61, 63, 65 nationality, 209 Native Americans, 181, 182, 184 native species, 136 natural resource management, 4, 11, 34, 39, 43, 48, 49, 50, 52, 54, 57, 59, 62, 64, 66, 100, 106, 107, 108, 109, 121, 158 natural resources, vii, viii, 1, 3, 6, 7, 10, 13, 20, 23, 24, 25, 27, 30, 31, 35, 43, 47, 48, 53, 54, 55, 56, 58, 59, 63, 66, 73, 88, 100, 106, 120, 124, 157, 158, 170, 171, 172, 206, 207, 215, 228 natural science, 36 nature conservation, 157 negative attitudes, 35, 108, 118, 119, 123, 124 neglect, 13, 22 negotiating, 110 negotiation, 56, 110 Nelson Mandela, 155, 177 Nepal, v, vii, viii, 105, 107, 108, 110, 111, 112, 113, 117, 118, 123, 125, 126, 127, 128, 206, 216 Netherlands, 36, 38, 40, 209 neutral, 77, 84 New Deal, 189 NGOs, viii, 6, 21, 26, 31, 34, 39, 105, 108, 114, 115, 116, 119, 120, 122, 123, 124 nickel, 30 Nigeria, 7, 14, 15, 16, 17, 20, 29, 38 Nile, 56 nitrates, 227 N-N, 148 non-renewable resources, 172 North America, 172, 175, 198 novelty seeking, 206, 217 NPS, 2, 191 nuisance, 10 nutrient, 171

O obstacles, vii, 1, 6, 187 OECD, 106, 127 offenders, 57, 76, 77, 82, 85 officials, 8, 23, 29, 35, 82, 90, 93, 94, 97, 110, 115, 116, 118, 123 oil, 7, 20, 28, 29, 30, 191 Oklahoma, 189 one dimension, 18 opaque governance, 34 openness, 123 operations, 31, 89, 90, 94, 95, 98, 172, 184, 187, 189

opportunism, 153 opportunities, x, 9, 35, 69, 99, 120, 150, 160, 172, 181, 187, 193, 194, 202, 204, 206, 207, 210, 215 optimists, 17 organic matter, 224 oryx, 171 outreach, 34 overgrazing, 188 overharvesting, 59 overlap, ix, 32, 130, 131, 148 overlay, 163 oversight, 25, 94, 95, 96, 97 ownership, 6, 7, 8, 9, 21, 32, 33, 52, 56, 61, 64, 70, 74, 86, 87, 89, 90, 93, 94, 96, 98, 99, 100, 149, 158, 172, 173, 185, 194

P Pacific, 2, 175, 189 palm oil, 7, 28 parallel, 65 participant observation, 6, 35 participants, 25, 60, 90, 93, 116, 122, 123 pastures, 157 PCP, 114, 115 peace, 185, 196 penalties, 55, 57, 59, 106 performance measurement, 109 permission, 12, 121 permit, 54, 77 petroleum, 13 pH, 227 pharmaceutical, 36 phenotype, 149 Philippines, 50 physical environment, 27 physical features, 109 pigs, 29 pilot study, 208, 209 pipeline, 27, 28, 30 plants, 13, 17, 25, 26, 36, 50, 55, 68, 71, 102, 103, 118, 120, 121, 161, 184, 225 platform, 109 pleasure, 216, 217 Pliocene, 222, 230 PM, 11, 19 poison, 12 policy, vii, 2, 4, 7, 24, 33, 42, 54, 56, 58, 61, 62, 63, 64, 65, 66, 87, 97, 99, 100, 101, 108, 118, 120, 130, 154, 157, 158, 160, 199 policy instruments, 63, 64 policy makers, 33, 66 policy reform, 160

Complimentary Contributor Copy

Index policymakers, 4 political instability, 58 political leaders, 66, 116, 193 political power, 31 political system, 8 political uncertainty, 58 politics, 94 pollution, 30, 210, 215, 228 polygamy, 187 ponds, 30, 224, 226, 227, 230 population, 5, 10, 12, 13, 16, 17, 19, 22, 27, 28, 30, 36, 64, 66, 73, 87, 120, 156, 209 population density, 17, 27, 87 population growth, 5, 13, 64 portraits, 185 Portugal, vii, x, 201, 202, 207, 208, 214, 219 positive attitudes, viii, 105 potato, 88 poultry, 89 poverty, vii, ix, 4, 5, 10, 11, 16, 21, 23, 25, 30, 34, 48, 50-54, 59, 61, 62, 63, 64, 65, 66, 69, 74, 86, 87, 90, 98, 99, 100, 102, 155, 158, 160, 171 poverty alleviation, vii, 4, 11, 23, 25, 30, 48, 51, 52, 53, 54, 59, 61, 62, 63, 64, 65, 66, 69, 74, 86, 90, 99, 100, 160 poverty eradication, 48, 61, 62, 63, 64, 66 power relations, 99, 100, 157 precedent, 11 precipitation, 226 preparation, 110, 125 preservation, 8, 25, 57, 130, 157, 173, 194 President, ix, 58, 179, 180, 183, 192, 193, 194 President Clinton, 194 prestige, 203 principles, viii, 2, 4, 23, 33, 64, 86, 89, 90, 98, 99, 103, 108, 109, 124, 153 private sector, vii, 2, 33, 158, 173 privatization, 9 problem-solving, 111 professionals, 51 profit, 36 programming, 152 project, ix, 16, 18, 26, 27, 28, 29, 51, 88, 89, 90, 91, 92, 93, 94, 95, 97, 98, 99, 110, 114, 116, 118, 120, 122, 123, 124, 150, 155, 158, 160 propagation, 12 property rights, 9 proposition, 174 protected areas, vii, x, 1, 2, 3, 4, 10, 11, 12, 18, 19, 25, 27, 29, 37, 42, 44, 48, 55, 57, 58, 62, 63, 64, 65, 66, 84, 87, 88, 89, 106, 108, 113, 117, 119, 120, 124, 151, 158, 173, 180, 201, 202, 205, 206, 207, 208, 210, 214, 217

241

protection, vii, ix, 1, 4, 7, 9, 12, 16, 22, 23, 24, 29, 33, 57, 58, 62, 63, 65, 70, 109, 110, 149, 171, 179, 183, 188, 191, 192, 193, 220, 228 protectionism, vii, 2, 58, 158 psychological pain, 26 psychological states, 204 public domain, 193 public officials, 35 public radio, 94 public resources, 199 pulp, 36

Q quality of service, 204 quantification, 119, 130, 131 quartz, 224 query, 163 questioning, 194 questionnaire, 115, 116, 208, 209 quota sampling, 209 quotas, 71

R radiation, 161, 163 radius, 30 rain forest, 9, 41, 102 rainfall, 14, 17 rainforest, vii, 1, 14, 45 rangeland, 109, 156, 198 rate of change, 146 rationality, 117 RDP, 228 real property, 33 reality, 11, 12, 20, 22, 23, 32, 33, 121, 122, 157, 190 reallocation of resources, 33 reciprocity, 39 recognition, 32, 33, 57, 100, 158, 173 recommendations, 6, 37, 98, 102, 152 recreation, ix, 179, 180, 189, 192, 193, 204, 207, 208, 217 recreational, 9, 193, 206 recycling, 171 redistribution, 8 reform(s), 13, 32, 34, 42, 65 regenerate, 120, 122 regeneration, 34, 122, 220 regression, 78, 203 regression analysis, 78 regulations, 23, 35, 50, 54, 55, 59, 107, 113, 118 relaxation, 203, 206, 215

Complimentary Contributor Copy

Index

242

relevance, 118, 203 relief, 188, 189 religion, 40, 181, 206 remittances, 33 remote sensing, 160 rent, 28 replication, 10 representativeness, 31, 130 requirements, 49, 153, 165, 166, 167, 168 RES, 62 researchers, x, 51, 88, 201, 203, 207, 209 resentment, 43, 75, 87 reserves, vii, 1, 3, 7, 9, 10, 12, 16, 21, 22, 37, 56, 57, 58, 61, 64, 113, 130, 152, 153, 156, 157, 158, 160, 172, 173, 176 resettlement, 20, 27 Residential, 113 residues, 228 resilience, 27, 32, 109 resolution, 40, 41, 109, 110, 131, 161, 162, 190 resource management, 4, 21, 25, 49, 50, 56, 59, 61, 100, 106, 107, 109, 121, 157, 159 response, 30, 49, 53, 59, 62, 69, 74-81, 85, 86, 97, 119, 152, 159, 188, 192, 193, 227 restaurants, 203, 210 restoration, 164, 193 restrictions, 21, 27, 194 restructuring, 33 retirement, 194 revenue, 11, 21, 50, 53, 57, 61, 64, 65, 69, 101, 106, 113, 115, 117, 160 rhetoric, 120 rights, vii, 1, 6, 7, 8, 11, 14, 16, 19, 21, 27, 30, 31, 32, 33, 40, 65, 110, 121, 157, 158, 172, 199 risk(s), 20, 22, 28, 41, 94 Romanticism, 189 root(s), 34, 49, 85, 156, 173 routes, 28, 164 Royal Society, 41, 151 rubber, 7 rule of law, 48 rules, 8, 9, 21, 31, 32, 54, 55, 59, 106, 107, 110, 111, 113, 117, 118, 120, 121, 122, 124, 142, 198, 207 rural areas, 13, 158, 160 rural development, 31, 49, 89, 98, 99, 180 rural people, 4, 158 rural poverty, 21, 34, 66, 158 Rwanda, 86, 101, 103

S safety, 28, 60, 205, 206, 207, 210 sanctions, 9, 55, 65, 107

sanctuaries, 3 savings, viii, 98, 105, 116 scarce resources, 4 school, 23, 49, 69, 90, 98 science, 22, 152, 154 scope, 32, 87, 110 sea level, x, 168, 219, 224, 228 sea-level, 231 seasonality, 167 secondary data, 206 security, 28, 32, 59, 172, 173 sediment(s), 222, 223, 224 segregation, 138 self-esteem, 206 self-sufficiency, 188 seller, 68 sensation, 203 sensitivity, 149 sensitization, 35, 172 services, viii, 2, 13, 33, 89, 130, 193, 203, 204, 205, 207, 208, 209, 210, 211, 212, 214, 215, 228 settlements, 17, 30, 55, 64, 184, 185 shade, 168 shape, 27, 31, 133, 148, 149 sheep, 157, 184, 187, 188 shelter, 7, 168 shores, 183 showing, 161, 163 shrubs, 14 significance level, 211, 213 simulation, 40 Singapore, 204 slaves, 182 small businesses, 28 small communities, 180 smuggling, 68 social capital, viii, 105, 109, 111, 118, 124 social change, 183, 189 social class, 116 social costs, 4 social development, 10, 11, 62 social environment, 11 social fabric, 193 social institutions, 33 social interactions, 107 social justice, 48 social learning, 121 social network, 23, 107 social norms, 107 social organization, 185 social rules, 107 social security, 28, 189 social services, 49

Complimentary Contributor Copy

Index social stress, 35 social structure, 59 social welfare, 171, 189 socialization, 203 society, 4, 55, 56, 86, 107, 118, 194, 196 soil erosion, 156 solitude, 196, 197 solution, viii, 35, 50, 129, 130, 150 South Africa, ix, 33, 152, 155, 156, 157, 158, 160, 172, 173, 174, 175, 176, 177, 205, 216 South America, 37 South Korea, 204 Spain, vi, vii, x, 152, 204, 209, 219, 220, 222, 229, 230 specialists, 68 species, ix, x, 3, 5, 7, 12, 14, 16, 17, 21, 27, 29, 30, 50, 55, 57, 68, 69, 120, 121, 130, 131, 136, 137, 144, 148, 149, 151, 152, 153, 155, 156, 157, 159, 160, 163, 164, 165, 166, 167, 168, 169, 172, 173, 208, 219, 220, 225, 228 species richness, 16, 130, 136, 144, 151, 153 speech, 194 Spring, 183, 191 stakeholders, 6, 10, 13, 20, 24, 25, 26, 33, 48, 49, 50, 66, 89, 90, 94, 95, 98, 99, 114, 124, 150 stars, 189 state(s), ix, 6, 7, 8, 9, 10, 11, 12, 14, 20, 22, 23, 26, 27, 30, 31, 32, 36, 44, 122, 124, 155, 156, 158, 173, 183, 185, 188, 192, 193, 203, 207, 217 state control, 158 state intervention, 7 state laws, 26, 31, 32 steel, 21 stereotypes, 157 stock, 157, 167 storms, 222, 230 stoves, 88 strategic planning, 158 stratification, 227 stress, 14, 35, 124, 204 structural adjustment, 13 structure, 24, 94, 97, 107, 117, 159 sub-Saharan Africa, 42 subsistence, 12, 18, 26, 30, 68, 69, 70, 113 substitutes, 89 substitution, 69 success rate, 20 Sudan, 33 Sun, 174, 205, 216 supernatural, 55 suppression, 33 surface area, 14 surplus, 68

243

surveillance, 16, 29 survival, 14, 23, 36, 120, 180 sustainability, viii, x, 2, 6, 13, 26, 33, 41, 89, 94, 98, 102, 109, 111, 156, 170, 173, 201 sustainable development, i, iii, vii, x, 4, 21, 34, 42, 43, 62, 170, 176, 216, 219, 228 Sweden, 1, 42 Switzerland, 37, 44, 102, 127, 176 symbolic meanings, 25 symmetry, 146 synthesis, 38

T Taiwan, 128, 206, 216 tanks, 88 Tanzania, 22, 33, 39, 40, 42, 43, 59, 175 tar, 191 target, 12, 51, 54, 97, 218 tax breaks, 106 taxa, 154 taxes, 27 TCR, 160, 161, 162, 164, 170, 173 teachers, 116 technical support, 109, 115 technology, 62 telecommunications, 193 tension(s), 117, 118, 203 tenure, 6, 8, 9, 23, 32, 33, 160, 172, 173 territorial, 228 territory, 3, 4, 11, 182 Thailand, 151, 204 threats, x, 30, 38, 49, 59, 69, 74, 153, 160, 219, 220 timber production, 56 time constraints, 204, 205, 207, 210 top-down, vii, 1, 25, 30, 117, 118 tourism, viii, ix, x, 2, 4, 13, 19, 33, 40, 50, 53, 63, 70-73, 79, 85, 103, 130, 132, 136, 149, 156, 172, 179, 189, 191, 193, 201-212, 214-219 tourism attractions, x, 201, 203, 205, 207, 208, 209, 210, 211, 212, 214, 215 trace elements, 230 tracks, 29, 30, 65 trade, 29, 68, 69, 98, 172, 182 trade-off, 172 traditional authority(s), 8, 26 traditions, 48, 107, 182, 194, 196 trafficking, 29 training, 16, 88, 114 transaction costs, 108, 109 transformation(s), 8, 9, 45, 57, 160, 167, 230 transgression, 224 translation, 35

Complimentary Contributor Copy

Index

244 transnational corporations, 36 transparency, 13, 48, 89, 95, 97, 109, 111, 123, 124 transport, 25, 29, 68, 222 transportation, 204, 206 treaties, 4 treatment, 122 trial, 76, 80, 82 triangulation, 185 tropical forests, 21, 49 tropical rain forests, 17 trust fund, 69 Trust Fund, 88, 89 Turkey, 151, 204

U UNESCO, 3, 17, 19, 29, 30, 36, 44, 102, 103, 220, 229 United Kingdom (UK), 3, 16, 33, 42, 44, 101. 125, 151, 154, 176, 209 United Nations (UN), 3, 4, 35, 37, 38, 44, 62, 102, 114 United Nations Development Programme (UNDP), 3, 5, 13, 44, 114 United States (USA), 3, 40, 102, 103, 128, 180, 185, 187, 194, 199 universe, 21 uranium, 191 urban, 8, 10, 136, 227, 230 urbanization, 10 Uzbekistan, 125

V validation, 24, 75 variables, 78, 109, 117, 160, 173 variations, 9, 133, 149, 222, 227 vegetables, 120, 121 vegetation, ix, 12, 37, 41, 111, 131, 149, 153, 155, 156, 159-163, 166-170, 173, 177, 227, 230 vehicles, 29, 85 vertebrates, 151, 225 victims, 36 vision, 21, 22, 124, 187, 194, 196 voluntary organizations, 31 vulnerability, 28, 106, 122, 136, 144, 149, 151

W Wales, 191 war, 122, 189 Washington, 37, 38, 39, 40, 43, 44, 102, 103, 125, 198, 199 waste, 188 water, 13, 23-25, 27, 69, 88, 91, 92, 108, 109, 115, 168, 180, 183, 185, 188, 189, 192, 206, 226-228 water policy, 109 water resources, 189 watershed, 7, 16, 185 waterways, 185 wealth, 61, 87, 116, 171 weapons, 17, 20 welfare, 183, 187 well-being, 20, 21, 22, 23, 88, 107, 108, 110, 111, 114, 115, 125 West Africa, 7, 38, 43, 128 wetlands, 3, 222, 224, 227, 228 wholesale, 10 wild animals, 57, 60, 61, 76, 79, 80, 82, 84, 85, 87, 88, 100 wilderness, 3, 192 wildlife conservation, ix, 22, 38, 42, 64, 65, 113, 114, 155, 156, 158, 159, 172, 173 wires, 123 Wisconsin, 39 wood, 13, 21, 30, 33, 55, 56, 68, 111, 113 wood products, 55 woodland, 168 workers, 25, 30, 188 workforce, 189 workload, 13 World Bank, 5, 13, 27, 28, 30, 37, 38, 40, 43, 44, 88, 90, 103, 109, 128, 160 World War I, 188, 189 WWW, 5

Y Yale University, 43 Yaounde, 36, 37, 41, 43

Z Zimbabwe, 158, 160, 176

Complimentary Contributor Copy