Sie sind auf Seite 1von 93

lessons for climate change adaptation from better management of rivers

Guest Editor: Jamie Pittock

VOLUME 1 ISSUE 3 2009

Lessons for climate change adaptation from better management of rivers

Climate and Development 1(3) November 2009. Published by Earthscan: Dunstan House, 14a St Cross Street, London EC1N 8XA, UK. # 2009 Earthscan All rights reserved. No part of this publication may be reproduced, stored in retrieval systems or transmitted in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, without written permission from the publisher. Cover image: Panos/Philippe Lissac ISSN: 1756-5529 (print), 1756-5537 (online) ISBN: 978-1-84971-090-9 Responsibility for statements made in the articles printed herein rests solely with the contributors. The views expressed by individual authors are not necessarily those of the editors, the funders or the publisher. SUBSCRIPTIONS Subscription prices for Volume 2: Institutions Online only: 304 $608 E456 Online & print: 320 $640 E480 (airmail extra) Personal Online only: Print only: 99 99 $199 E130 $199 E130 (airmail extra)

191

EDITORIAL Why a special issue on adaptation and water management?


JAMIE PITTOCK and STEPHEN DOVERS

194

PAPERS Lessons for climate change adaptation from better management of rivers
JAMIE PITTOCK

212

Floodplain restoration along the lower Danube: A climate change adaptation case study
SUZANNE EBERT, ORIETA HULEA and DAVID STROBEL

220

Freshwater management and climate change adaptation: Experiences from the Great Ruaha River catchment in Tanzania
JAPHET. J. KASHAIGILI, KOSSA RAJABU and PETRO MASOLWA

229

Adapting to climate change in the Godavari River basin of India by restoring traditional water storage systems
BIKSHAM GUJJA, SRABAN DALAI, HAJARA SHAIK and VINOD GOUD

241

Freshwater management and climate change adaptation: Experiences from the Central Yangtze in China
XIUBO YU, LUGUANG JIANG, LIFENG LI, JINXIN WANG, LIMIN WANG, GANG LEI and JAMIE PITTOCK

Orders can be placed online at www.earthscan.co.uk/ journals/cdev or sent to the journals distributors, Portland Customer Services, using the contact details below. Post: Portland Customer Services, Commerce Way, Colchester, CO2 8HP, UK Fax: 44 (0)1206 799331 Tel: 44 (0)1206 796351 Email: sales@portland-services.com Printed in the UK by MPG Books Ltd on FSC certied paper. The editors and publishers gratefully acknowledge support from the Swedish International Development Cooperation Agency (Sida) for Climate and Development Climate and Development is published quarterly. Periodicals Postage Paid at Rahway, NJ. US agent: Mercury International, 365 Blair Road, Avenel, NJ 07001. POSTMASTER: Address changes to CLIMATE AND DEVELOPMENT, 365 Blair Road, Avenel, NJ 07001.

249

Integrated river basin management in the Conchos River basin, Mexico: A case study of freshwater climate change adaptation
J. EUGENIO BARRIOS, J. ALFREDO RODRIGUEZ-PINEDA and MAURICIO DE LA MAZA BENIGNOS

261

Participatory river basin management in the Sao Joao River, Brazil: A basis for climate change adaptation?
LUIZ FIRMINO MARTINS PEREIRA, SAMUEL BARRETO and JAMIE PITTOCK

269

Embracing uncertainty in freshwater climate change adaptation: A natural history approach


JOHN H. MATTHEWS and A. J. WICKEL

www.earthscan.co.uk

Aims and Scope


Climate and Development is dedicated to the range of issues that arise when climate variability, climate change and climate policy are considered along with development needs, impacts and priorities. It aims to make complex analysis of climate and development issues accessible to a wide audience of researchers, policymakers and practitioners, and to facilitate debate between the diverse constituencies active in these elds throughout the world. The journal provides a forum to communicate research, review and discussion on the interfaces between climate, development, policy and practice. Every three months it presents conceptual, policy-analytical and empirical studies of the interactions between climate impacts, mitigation, adaptation and development on scales from the local to global. Contributions from and about developing countries are particularly encouraged; however, research on developed countries is welcome provided that the link between climate and development is the central theme. Climate and Development is of direct and vital relevance to academics, policy analysts, consultants, negotiators, industrial and non-governmental organisations, and to all those working to ensure a better understanding of the links between climate and development. The journal is the platform of choice for academic debate on issues that link climate and development, and invites contributions on all such issues. These include, but are not limited to:
B B B B B B B B B B B B B

The vulnerability of communities to the combined impacts of climate change and non-climatic stresses Links between development and building capacity to respond to climate change The integration (mainstreaming) of climate policy adaptation and mitigation into sectoral planning and development policy Conicts and synergies between mitigation, energy development and poverty The importance of climate and long-term weather forecasting for development Responsibilities of developing countries in a post-2012 climate policy regime The effects of climate change on meeting the Millennium Development Goals The implications for development of the UN Framework Convention on Climate Change and its Kyoto Protocol, as well as all other existing or proposed policy frameworks Financing arrangements for adaptation and mitigation in developing countries Economic analysis of the effects of climate adaptation and mitigation on developing countries Traditional knowledge and local strategies for managing natural resources and coping with climate change Forest management and its relationship to mitigation, adaptation and development Adaptation, mitigation and the poor

These and other topics are addressed in a number of ways, including:


B B B B B B

Research articles (theoretical developments, concepts and methods, empirical analysis and policy assessments) Review articles Case studies Viewpoints Book reviews Meeting reports

editorial

Why a special issue on adaptation and water management?


JAMIE PITTOCK* and STEPHEN DOVERS
Fenner School of Environment & Society, Australian National University, Canberra ACT 0200, Australia

The Copenhagen Climate Change conference in March 2009 reported that adaptation measures to lessen the impacts of climate change are urgently needed now. Given the considerable uncertainties around projections of climate impacts on water resources at local and regional scales, building resilience, managing risks, and employing adaptive management are likely to be the most effective adaptation strategies (Richardson et al., 2009). The conference went on to conclude: As part of building effective adaptation, research is urgently required into the implications of existing policies and potential future policies with regard to adaptation: do they support or hinder adaptation, and how do they need to be changed? (Richardson et al., 2009). This special edition of Climate and Development aims to contribute to the ongoing process of learning how our societies may more effectively adapt to a changing climate. We start here by outlining the choice of focus of this volume and summarizing the papers that comprise it. We conclude by highlighting the key lessons drawn from this research.

cope with climate change, may lessen the consequences of that change. Such adaptations are widespread in the water sector, although with varying degrees of effectiveness in coping with climate change . . . (Bates et al., 2008). As such autonomous adaptations are widespread and possibly the most common form of adaptation to climate change, there is much that society can learn from the factors that hinder and facilitate the effectiveness of such measures, and from understanding learning processes and the limits of adaptation: this is the focus of this special edition.

2. Case studies
Climate and Development was established (in part) to: make complex analysis of climate and development issues accessible to a wide audience of researchers, policymakers and practitioners, and to facilitate debate between the diverse constituencies active in these elds throughout the world, and to offer a possibility of publication for many of the practical lessons that are learnt in projects but often not shared with the academic community. This special edition lls such a role by reporting on the lessons drawn from six empirical, consistently designed freshwater adaptation case studies from developing countries, based on projects of the conservation organization WWF (World Wildlife Fund/World Wide Fund for Nature). These case studies illustrate a number of issues at the forefront of the global debates on sustainable water management and climate change adaptation: B Gujja et al. report on their work in India that assesses the costs and benets from restoring traditional village water tanks as an adaption measure. Pittock then

1. Focus
We chose to focus this edition of Climate and Development on water management because it is an important eld from which to draw lessons on risk management and adaptation. The Intergovernmental Panel on Climate Change (IPCC) declares that adaptation to changing conditions in water availability and demand has always been at the core of water management (Kundzewicz et al., 2007). The IPCC dene such historical actions in the water sector as autonomous adaptation measures that do not constitute a conscious response to climate stimuli, but result from changes to meet altered demands, objectives and expectations which, whilst not deliberately designed to

B *Corresponding author. E-mail: jamie.pittock@anu.edu.au


CLIMATE AND DEVELOPMENT 1 (2009) 191193 doi:10.3763/cdev.2009.0019 # 2009 Earthscan ISSN: 1756-5529 (print), 1756-5537 (online) www.earthscanjournals.com

192 Pittock and Dovers

compares this approach with the proposed construction of a large dam on the Godavari River, encapsulating the debate between proponents of adaptation through large infrastructure vs. decentralized and small-scale appropriate technologies. Yu et al. outline the benets of restoring ood plain lakes in the central Yangtze River basin compared to reliance on ood protection dykes. This paper also highlights the enhanced livelihoods derived from more ood-adapted agri- and aqua-cultural systems, and the importance of concurrent interventions at different geopolitical scales within China. Barrios et al. detail their work on enhancing water security in the drought-prone, over-allocated Rio Conchos basin in northern Mexico. Their paper illustrates the need for conjunctive management of surface and ground waters, the need to use multi-stakeholder processes to better manage scarce resources, and the opportunities for international treaties to drive local reforms. Pereira et al. outline the development of adaptive management capacities through a multi-stakeholder river basin consortium at Rio Sao Joao in Brazil. They high light the enabling power of sound national water law in facilitating basin-scale institution building, community engagement and adaptive management to progressively address environmental problems. Ebert et al. describe reform of river management in the lower Danube basin in Eastern Europe. In outlining the adaptation benets from large-scale oodplain restoration for ood management, economic diversication and biodiversity conservation, they demonstrate how supranational European institutions have driven reform. Kashaigili et al. detail outputs from a programme to restore dry season ows in the Great Ruaha River of Tanzania. They illustrate the benets of concurrent institutional interventions to reduce poverty and direct interventions to reduce vulnerability to water scarcity.

places. Yet by drawing on multiple cases and using consistent analytical frameworks they usefully report the trials and errors and successes of social learning in multistakeholder adaptive management processes (Lee, 1993). The case studies highlight uncertainties in these societies responses to water management and climate variability and change. Yet it is precisely these sorts of real-life examples of adaptation in uncertain conditions that researchers, policymakers and society at large need to learn from if the global community is to better adapt to the problems of climate change and water management that afict our globe. WWFs willingness to expose its work to academic scrutiny is to be welcomed and, consequently, it deserves to be rewarded in terms of constructive engagement to further enhance their programmes. It is to be hoped that this volume inspires other practitioner organizations to publish similar assessments of their programme portfolios.

3. Key lessons
In this volume, Pittock reviews the six case studies to derive common lessons on policy style and sub-programme detail (Dovers, 2005), to inform practitioners, policymakers and our broader societies on measures that may enhance adaptation to climate change. Crucially, he observes that a number of charismatic local programme leaders had not engaged in climate adaptation, concluding that adaptation proponents need to engage better with local institutions (Burgess et al., 1998; Meinke et al., 2006) to seize the opportunities for complementary no and low regrets adaptations in their current activities. The assessment of these freshwater cases in developing countries contributes to debates in the literature, by proposing that climate change adaptation is best enhanced by: B concurrently acting to reduce poverty and enhance livelihoods, and manage biophysical vulnerability, rather than favouring either response alone (Adger, 2006; Schipper, 2007); B favouring investment in scalable, decentralized, small-scale appropriate technologies, and enhancing environmental resilience (Tompkins and Adger, 2004), rather than rst opting for centralized infrastructure; B investing in the capacity of local- to basin-scale institutions to apply adaptive management programmes over many years (Connor and Dovers, 2004); B linking institutions at different geopolitical scales to facilitate better local to global adaptation (Adger

These case studies from practitioners do not fully conform to the idealized formulas of academic research. However, given uncertainties associated with climate change impacts and the urgent need to distil and communicate lessons for adaptation in the near term, there is great value in examining programmes that have (in all but one case) been operating for more than ve years. These retrospective studies of autonomous adaptation in projects by WWF and its partners lack desirable quantitative data in

CLIMATE AND DEVELOPMENT

Editorial 193

et al., 2005), which in most cases will require more effective and efcient national institutions. Matthews and Wickel in this volume draw on the lessons derived from this WWF work. They note that climate change impacts on freshwater systems are associated with high uncertainty and criticize model-driven impacts thinking. Identifying the need for multi-generational response, they propose an adaptation thinking approach as a template for sustainable development and climate change adaptation. Dovers (2009) proposes that we can go at least halfway to a believable adaptation policy by implementing known, well-supported policy and management options and by normalizing adaptation, and empowering ofcials, agencies, local communities. Pittocks assessment identies that in many cases national governments have failed to turn policy into effective action, especially in terms of implementing enabling laws and nancing measures for sub-national adaptive management institutions, particularly river basin management organizations. These projects also highlight the extensive opportunities in the freshwater and climate adaptation eld for no and low regrets interventions: restoration of environmental resilience and other ecological services; scalable, decentralized, small-scale appropriate technologies; and effective, multi-stakeholder adaptive management institutions. These case studies expose the limits of expert- and modelling-driven adaptation methods by showing that knowledgeable and well-meaning local leaders may postpone action while awaiting better advice and data, when the climate impact uncertainties are unlikely to be reduced to a meaningful extent any time soon. The research also emphasizes the tremendous opportunities available to implement practical adaptation measures now.

References
Adger, W. N., 2006. Vulnerability. Global Environmental Change, 16(3). 268 281. Adger, W. N., Arnell, N. W. and Tompkins, E., 2005. Successful adaptation to climate change across scales. Global Environmental Change Part A, 15(2). 7786.

Bates, B. C., Kundzewicz, Z. W., Wu, S. and Palutikof, J. P . (eds). 2008. Climate Change and Water. Technical Paper of the Intergovernmental Panel on Climate Change. IPCC Secretariat, Geneva. Burgess, J., Harrison, C. M. and Filius, P 1998. Environ., mental communication and the cultural politics of environmental citizenship. Environment and Planning A, 30. 1445 1460. Connor, R. and Dovers, S., 2004. Institutional Change for Sustainable Development. Edward Elgar Publishing, Cheltenham, UK and Northampton, USA. Dovers, S., 2005. Environment and Sustainability Policy: Creation, Implementation, Evaluation. Federation Press, Annandale, VA. Dovers, S., 2009. Normalizing adaptation. Global Environmental Change, 19(1). 4 6. Kundzewicz, Z. W., Mata, L. J., Arnell, N. W., Doll, P Kabat, ., P Jimenez, B., Miller, K. A., Oki, T., Sen, Z. and ., Shiklomanov, I. A., 2007. Freshwater resources and their management. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, M. L. Parry, O. F. Canziani, J. P Palutikof, P J. van der Linden and C. E. . . Hanson (eds). Cambridge University Press, Cambridge, UK. 196. Lee, K. N., 1993. Compass and gyroscope: integrating science and politics for the environment. Island Press, Washington, DC and Covelo, CA. Meinke, H., Nelson, R., Kokic, P Stone, R., Selvaraju, R. ., and Baethgen, W., 2006. Actionable, climate knowledge: from analysis to synthesis. Climate Research, 33. 101 110. Richardson, K., Steffen, W., Schellnhuber, H. J., Alcamo, J., Barker, T., Kammen, D. M., Leemans, R., Liverman, D., Munasinghe, M., Osman-Elasha, B., Stern, N. and Waever, O., 2009. Synthesis Report. Climate Change. Global Risks, Challenges and Decisions. University of Copenhagen, Copenhagen. Schipper, E. L. F., 2007. Climate Change Adaptation and Development: Exploring the Linkages. Tyndall Centre Working Paper No. 107. Tyndall Centre for Climate Change Research, Norwich, UK. Tompkins, E. L. and Adger, W. N., 2004. Does adaptive management of natural resources enhance resilience to climate change? Ecology and Society, 9(2). 10.

CLIMATE AND DEVELOPMENT

review article

Lessons for climate change adaptation from better management of rivers


JAMIE PITTOCK*
Fenner School of Environment & Society, Australian National University, Canberra ACT 0200, Australia

Autonomous adaptation in the water sector is assessed to derive lessons for more successful climate change adaptation from six empirical, consistently designed river management case studies based on projects of WWF. They show that when adaptation measures are considered in the context of common problems in water management, many practical ways of building resilience to climate change through mainstream programs are evident. The cases are mainly from developing countries India, China, Mexico, Brazil, the lower Danube basin and Tanzania where efforts to reduce environmental degradation and enhance livelihoods have directly helped to reduce vulnerability to natural hazards and climate change. The key lessons include: the benets of concurrent measures for improving livelihoods and reducing physical vulnerability; the need to enhance and fund local institutions to mainstream adaptation programmes; and the value in implementing no and low regrets measures despite uncertainties. Keywords: adaptation; climate change; developing countries; institutions; non-governmental organizations; rivers; water

1. Introduction
The world faces grave challenges in sustaining water resources for people and nature, problems that are exacerbated by the impacts of climate change and the need for ongoing, effective and efcient adaptation. The term adaptation can be broadly applied to actions to manage changes in the environment or society, beyond impacts induced by climate change. The Intergovernmental Panel on Climate Change (IPCC) declares (Kundzewicz et al., 2007, p. 196) that: Adaptation to changing conditions in water availability and demand has always been at the core of water management. The IPCC also denes autonomous adaptation actions as (Bates et al., 2008, p. 48): those that do not constitute a conscious response to climate stimuli, but result from changes to meet altered demands, objectives and expectations which, whilst not deliberately designed to cope with climate change, may lessen the consequences of that change.
B *E-mail: jamie.pittock@anu.edu.au
CLIMATE AND DEVELOPMENT 1 (2009) 194211 doi:10.3763/cdev.2009.0021 # 2009 Earthscan ISSN: 1756-5529 (print), 1756-5537 (online) www.earthscanjournals.com

Such adaptations are widespread in the water sector, although with varying degrees of effectiveness in coping with climate change. As autonomous adaptations are widespread and possibly the most common form of adaptation to climate change, there is much that society can learn from the factors hindering and facilitating the effectiveness of such measures, and this is the focus of this paper. Further, as the climate will continue to change, adaptation is considered in this paper to be an ongoing rather than nite process (Matthews and Wickel, 2009). To contribute to the design of more effective freshwater climate adaptation processes, this paper considers freshwater case studies that meet the IPCCs denition of autonomous adaptation to derive lessons on what motivated these societies to change, the factors that led to more successful processes, and how interventions may best be sustained. Rather than a theoretical assessment of what measures could or should be

Lessons for climate change adaptation from better management of rivers 195

implemented, this paper seeks lessons from ongoing freshwater adaptation processes. The paper also considers the benets of these autonomous adaptation measures in terms of how they increase resilience1 (Bates et al., 2008), and reduce vulnerability2 (Bates et al., 2008). In many cases climate change is expected to be felt rst, and most severely, by changes in hydrology. In response, development of effective policies requires in part practical implementation knowledge as one key evidence base (Head, 2008), as well as scientic and political knowledge. While these case studies have elements of all three types of knowledge, it is lessons from ongoing implementation that are sought in this paper. Rather than a search for a complete package of programme elements (Dovers, 2005), the comparative policy analysis undertaken in this research is focused on subprogramme detail, derivation of lessons from specic elements of the processes, and also the policy style. In 2008, in presenting the preliminary ndings for UN Waters 3rd World Water Development Report, the reports content coordinator, Dr William J. Cosgrove, regretted the lack of published case studies that linked freshwater management and its potential to contribute to climate change adaptation. He called on implementing agencies to publish assessments of their activities. In response, this paper is intended to identify such knowledge from the work of a large non-governmental organization. This paper reports on the global lessons drawn from six empirical, consistently designed case studies of autonomous freshwater adaptation processes based on projects of a conservation organization, the World Wide Fund for Nature (also known as the World Wildlife Fund or WWF). The cases are from India (Gujja et al., 2009), China (Yu et al., 2009), Mexico (Barrios et al., 2009), Brazil (Pereira et al., 2009), the lower Danube basin (Ebert et al., 2009) and Tanzania (Kashaigili et al., 2009). Project sites were selected by the largely independent local WWF ofces at different times, although three of the six projects were substantially funded in

the period reported on here through a globally coordinated programme called Investing in Nature, supported by the Hong Kong Shanghai Banking Corporation (HSBC) and WWF UK. The six river basins concerned were all considered by WWF to be signicant for biodiversity conservation, and their conservation work commenced more than six years ago at all sites, except with the Godavari project. Otherwise, the only common thread in their selection was a need perceived by WWF and sectors of the local community to respond to severe environmental degradation, often indicated by disasters, which threatened biodiversity and peoples livelihoods (Table 1). In response to this environmental degradation, WWF and the local institutions instigated actions that reduce vulnerability to climate variability and related natural resource management problems, including the types of climate impacts expected to be exacerbated by climate change. WWF is a proponent of the sustainable development environmental discourse (Dryzek, 1997), and its actions in these eld projects reect their beliefs in nested social and ecological systems, that environmental protection and socioeconomic benets are mutually reinforcing, and in decentralized, exploratory and variable approaches in pursuit of sustainability (Lee, 1993). A key dilemma facing policymakers is whether adaptation is better facilitated by focusing on biophysical risk reduction, or whether it would be more effective to invest in reducing poverty and improving livelihoods more generally so as to build the resilience and adaptability of local communities to climate change impacts (Brooks, 2003; Adger, 2006; Schipper, 2007). The research considers how such measures are best integrated into society (Ross and Dovers, 2008). A further choice is between more technical infrastructure on the one hand, and on the other, favouring small-scale and decentralized interventions with a greater emphasis on increasing societal capacities (Moench and Stapleton, 2007; Ribot et al., 2009). These questions are further assessed in this paper.

CLIMATE AND DEVELOPMENT

196 Pittock

TABLE 1 Environmental degradation and disasters that instigated WWF and societal responses Basin and location Major environmental degradation and disasters Maner River tributary of the Godavari River, India Water scarcity an increasing problem in the region as populations increase and water resources are extensively exploited. Access to water is a focus of many government and community organizations programmes. Lakes in the central Yangtze River basin, China Floods, drought, pollution, shery decline all increasing problems. Major oods in 1995, 1996, 1998 and 1999 sparked responses from governments. 2002 to present Demonstrate that re-linking oodplain lakes to the Yangtze River, and promoting more diverse and ood-adapted livelihood activities would improve water quality, biodiversity conservation and the livelihoods of local people. Rio Conchos, Mexico Water scarcity drought from 1994 to 2006 instigated responses from stakeholders. 2002 to present Improve the condition of freshwater ecosystems in the Rio Grande/Bravo basin by promoting the application of integrated river basin management. Rio Sao Joao (Rio de Janeiro State), Brazil Pollution, shery decline, water scarcity. By 1999 eutrophication of water bodies had largely eliminated the inland sheries, and reduced water access and sparked community demands for rectication. Lower Danube River, Romania, Bulgaria, Moldova and the Ukraine Great Ruaha River, Tanzania Floods, pollution both increasing problems. Major oods in 19982002, 2005 and 2006 resulted in demands for more effective management by governments and communities. Water scarcity river ceased owing in the dry season from 1993 resulting in a 2001 Prime Ministerial commitment to restore river ows. 2003 to present 1992 to present Establishment of the Danube River basin as a model of nature conservation and community prosperity, including restoration of freshwater and forestry resources along the lower Danube. To enable the people of the Great Ruaha River catchment to plan, manage and utilize their water and related natural resources sustainably, and by doing this, alleviate poverty and improve livelihoods. 1999 to present Restore the water quality and biodiversity of water bodies in the Sao Joao region by promoting the application of integrated river basin management. WWF project period March 2005 February 2007 WWF project objective/s (as summarized by the author) Assess the socio-economic and environmental costs and benets of restoring traditional village water tanks as an alternative to major infrastructure schemes to increase water supplies.

2. Methods
Six existing WWF projects (see Table 1) were selected for research by the author in consultation with staff of WWF UK. The projects were selected on the basis that they had the following characteristics: a signicant focus on peoples management of hydrological variability; were from countries with developing or emerging

economies and from a broad continental distribution; and had been under way for sufcient time to have produced substantial outputs. This assessment was undertaken between February and December 2008. Each WWF project was funded to employ a local consultant reporting to the local WWF ofce to prepare a case study report responding to an analytical framework. The reports covered the background to

CLIMATE AND DEVELOPMENT

Lessons for climate change adaptation from better management of rivers 197

the work and the outputs and lessons in three areas: adaptation, livelihoods and conservation. They were prepared iteratively in consultation with the author in order to clarify data and increase consistency between the reports. The questions that were applied to each of the projects are detailed in the Annex. The case studies were then analysed by the author. The work of these projects involved social and institutional changes as much as or more than biophysical and technological interventions. The measures deployed in these autonomous adaptation processes can be categorized as:
B

B B

B B B

Decommissioning or changing the operations of underperforming infrastructure, like ood protection dykes and sluice gates. Restoring the ability of the natural environment to provide ecosystem services, such as oodwater retention, storing water in aquifers, water purication and sheries. Adopting locally available and small-scale technologies, such as village water tanks. Changing agricultural and aquacultural practices to more sustainable methods that: produce fewer pollutants; reuse water, such as for sh production then irrigation; are more water efcient; require less inputs; and secure higher returns for more valued produce. Providing better waste management systems, especially for sewerage. Diversifying local livelihoods into more profitable and less water-dependent enterprises. Increasing the incomes derived from natural commodities, such as sh, to reward producers adopting more sustainable practices and increase the resilience of these households. Establishing and strengthening local institutions to facilitate adaptive management and self-determination, including establishing and enforcing more sustainable behavioural norms for uses of natural resources such as water. Facilitating basin-scale multi-stakeholder institutions to: establish partnerships; develop common visions; lead adaptive management;

and connect the local to global measures needed for more effective adaptation and sustainability. B Advocating laws and government programmes that facilitate subsidiarity by providing basin and local institutions with the mandate and access to resources for adaptive management. B Improving connectivity in freshwater ecosystems by applying environmental ows, ensuring wildlife passage through or over water infrastructure, and restoring riparian habitats. B Restoring habitats to increase the resilience of these ecosystems to climate impacts, and their capacities to support greater populations of ora and fauna species, especially those that are threatened or of economic value.

3. Results
Table 2 summarizes the main adaptation, livelihood and conservation benets to date from the six projects. Successful outcomes to date from these autonomous adaptation cases can be categorized under the following:
B B B

Flood retention: increased capacity to safely retain higher peak ood ows. Water security: more reliable access to water in areas prone to scarcity. Pollution reduction: cuts to pollution levels and the risk that pollution impacts like eutrophication will be exacerbated by higher temperatures resulting from climate change. Livelihoods: diversied income generation strategies and increased incomes of many participants that may increase resilience of communities to climatic events. Institutional capacity: established and strengthened local institutions, increasing their adaptive management capacities. Connectivity: re-linked habitats and populations of species, enabling greater mobility and capacity to colonize new habitats that may be required to survive in a warmer world.

CLIMATE AND DEVELOPMENT

198 Pittock

TABLE 2 Summary of key climate adaptation, livelihood and conservation benets Project Likely major climate change impacts Lower Danube, eastern Europe Pollution exacerbated Biodiversity impacted Great Ruaha River, Tanzania Greater water scarcity Biodiversity impacted Water Users Associations and other basin institutions strengthened Godavari tanks, India Greater water scarcity Impacts of alternative adaptation options Programme adopted by the state government. Alternative to proposed USD$4 billion dam demonstrated Yangtze lakes, China Flooding increased Pollution exacerbated Biodiversity impacted Restored 450 km2 lakes. Can retain 285 mm of ood waters Pollution and the risk of algal blooms reduced Government adopted restoration policies Yangtze Forum established for adaptive management Continued
3

Key climate adaptation benets

Key livelihood benets

Key ecosystem benets

Increased ooding

Flood storage increased through restoration of oodplains. Plan to restore 2,250 km2. Of this area 14.4% has been or is being restored

Livelihoods diversied

Restored 4,430 ha of habitats and reconnected a 68 km2 lake to the river

Better access to clean water Pollution and the risk of algal blooms reduced Reduced vulnerability to drought Ecological services of EUR500/ha from restored oodplains Established 20 Community Banks Diversied into livelihoods with reduced reliance on water Greater surface and ground water access from restored tanks Increased agricultural production, employment and incomes Tank management systems established Reduced agricultural inputs

Fish and bird populations restored Protected areas expanded by 5,757 km2, including large areas of oodplains, in Romania Flows restored in some places

Water sources and riparian vegetation restored Tree felling for charcoal production reduced Enhanced habitats for birds in the tanks

Alternative to environmental damage from proposed new dam demonstrated

Cultural benets

Improved access to drinking water Fish resources increased Diversication of livelihoods and increased incomes

Restored 450 km2 lake habitats, new 60 km2 reserve Populations of sh, birds and Yangtze Porpoise increased

CLIMATE AND DEVELOPMENT

Lessons for climate change adaptation from better management of rivers 199

TABLE 2 Continued Project Likely major climate change impacts Rio Conchos, Mexico Greater water scarcity Biodiversity impacted Established institution for adaptive basin management Environment recognized as a user in the water law Rio Sao Joao, Brazil Biodiversity impacted Establishment of multi-stakeholder, adaptive, river basin management institutions Management approach adopted widely in other basins Pollution exacerbated Pollution cut by 75%, reducing algal blooms Increased economic efciency in agriculture Enhanced livelihoods of communities in the headwaters Restored 244 km2 coastal lagoons, rejuvenating tourism and shing industries Training and economic diversication Improved water supply Restored riparian, oodplain and lagoon habitats. Riparian corridors link remnant habitat of a threatened primate, the Golden Lion Tamarin River connectivity restoration planned Developing payment for ecological services and environmental ows Vulnerability to drought reduced More secure access to water Conservation of endemic sh Key climate adaptation benets Key livelihood benets Key ecosystem benets

Populations and habitats: restored populations of species and areas of habitat that may be better able to resist and survive impacts of severe climatic events.

4. Discussion
Reviewing the outputs from these six case studies, the following eight overarching lessons for more effective adaptation processes are identied and listed in Table 3, together with the most relevant examples. Considering these cases and lessons further, the following issues for effective adaptation are identied for wider discussion:

4.1. Quantiable targets


Like many organizations, WWF continually debates whether more targeted and sophisticated

programmes would achieve more benets for people and the environment. In these case studies, there is little doubt that more climate-informed and target-driven projects could achieve more effective interventions. For example, environmental ow methods are being applied globally to better dene the objectives of freshwater biodiversity conservation and the thresholds for the quantity and quality of water required to achieve them under the assumption of a stationary, natural hydrological regime. These methods could be applied to maintain specic freshwater biodiversity values under conditions of climate change (Anon., 2007). In the Ruaha and Rio Conchos projects, the generic interventions to attenuate water scarcity are buying both time and stakeholder ownership of the development of scientically based, quantitative environmental ows. This suggests that taking action to adapt to the most obvious problems should not wait for more precise information. By

CLIMATE AND DEVELOPMENT

200 Pittock

TABLE 3 Lessons derived from the six case studies Lesson derived 1. Local ownership. Participation of local stakeholders increased the sustainability and effectiveness of the measures. Supporting examples Ruaha: Community ownership through Water Users Associations has been essential to agree, implement and enforce measures beyond the governments reach; Godavari: Villagers contributed two-thirds of the resources needed for tank restoration and established local management institutions; Rio Conchos: The Inter-institutional Working Group and work with irrigators and villagers in the rivers headwaters have sustained major interventions; Rio Sao Joao: Establishment of the basin Consortium, including local governments, has engaged wide sections of the community and driven reforms to water management. Qualifying examples Danube: To some extent the national and international demand for better ood control has prevailed in place of local community ownership of restoration of the Danube oodplains, but the absence of local consent has delayed progress at a number of sites; Yangtze: While local people and national authorities both owned restoration of the oodplain lakes for improved environmental quality and livelihoods, local institutions do not appear to support the use of these lakes by national authorities for ood management purposes. 2. Immediate benets. Local stakeholder support depends on receipt of immediate benets; these appeared to engender support for more challenging measures. Ruaha: Establishment of Community Conservation Banks and other livelihood benets has underpinned support for environmental ow assessments; Yangtze: Enhanced livelihoods at initial sites has seen support for restoration of additional oodplain lakes; Rio Conchos: Initial benets from more efcient water use have enabled consideration of environmental ow allocations; Rio Sao Joao: Achievement of initial plans for reduced pollution and restored sheries has led to plans for new measures to restore riparian corridors and the watershed. 3. Multiple benets. Many freshwater adaptations to climate change impacts are practical now, can be scaled up, and had multiple environmental and socio-economic benets. Danube: Floodplain restoration offers immediate benets and could be incrementally scaled up to manage increased ood risks; Ruaha: Watershed restoration activities in half the districts in the basin delivered benets, including increased river ows, and could be expanded to other districts; Continued -

CLIMATE AND DEVELOPMENT

Lessons for climate change adaptation from better management of rivers 201

TABLE 3 Continued Lesson derived Supporting examples Godavari: Village tank restoration is a cheap option for increasing water supply and could be expanded to cover more than 200,000 such tanks in India; Yangtze: Restoration of 200 km2 oodplain lakes enhanced livelihoods, environmental quality and ood control, and could be expanded to a much larger oodplain area; Rio Conchos: Watershed restoration and water efciency measures have succeeded locally in reducing vulnerability to drought and could be scaled up considerably; Rio Sao Joao: The institutions established for water pollution reduction have succeeded on this problem and are now moving progressively to address other adaptation challenges in water and basin management. 4. Linking local to national to global. The most effective measures drew strength and linked institutions and action at different geopolitical scales. Danube: Obligations under the European Union and Danube Convention, national policies and local action have combined to initiate, fund and implement oodplain restoration; Yangtze: National policies for ood control and more sustainable water management enabled provincial and local government authorities to implement the oodplain lake restoration measures; Rio Conchos: Obligations under the border rivers treaty to deliver water to the USA and funding from the North American Development Bank, combined with the need within Mexico to reduce vulnerability to drought, resulted in effective water efciency measures; Rio Sao Joao: National and state water laws provided the mandate for the basin Consortium and underwrote its funding, which enabled local institutions to implement reforms more effectively. Continued Godavari: The case study of restoration of 12 village tanks involved relatively local-scale actions. Yet this case relied on international funding in a situation where state and national institutions had been ineffective in facilitating action. Scaling up application of tank restoration would require state and possibly national government support to succeed. Qualifying examples

CLIMATE AND DEVELOPMENT

202 Pittock

TABLE 3 Continued Lesson derived 5. Adaptive management. Effective adaptation was an iterative process over many years. Supporting examples Danube: Sequential adoption of stronger basin agreements, from the Danube Convention in 1994, Lower Danube Green Corridor Agreement in 2000 and the EU Water Framework Directive milestones from 2000 to 2015 have provided renewed impetus for adaptive management; Rio Sao Joao: Since its establishment in 1999 the basin Consortium has implemented three phases of measures, and as the preceding targets have been met, this has generated support for new interventions; Ruaha: To a lesser extent the sequence of national policies since 1991 and water projects in this basin are also an example of iterative processes enhancing awareness of the issues and options for responses over time. 6. Funding adaptation. Regular funding was needed to sustain adaptation. Godavari: Village institutions will collect water use fees to sustain management of restored tanks; Rio Sao Joao: The Consortium was funded through fees from municipal government members and local companies. As well, the head of the Consortium is seconded from the state government. These resources are used to leverage additional funding for management measures. Ruaha: Although promised in the national water policy the allocation of water use fees back to local management institutions has not occurred, jeopardizing the ongoing work of these organizations; Rio Conchos: State law frustrates efforts to establish binding payment for a watershed services scheme, which would enable urban dwellers to cross-subsidize watershed management and restoration measures. 7. Communicating adaptation. The language and perception of adaptation as new and complicated appeared to have stymied engagement of local communities and governments. Godavari, Yangtze and Rio Sao Joao: When rst discussed with programme leaders they expressed the view that the contradictory and uncertain scenarios from climate impact models for these places meant that it was not possible to dene climate change adaptation measures yet. Additional views expressed included: different opinions of Chinese academic and government ofcials on the nature of climate change; the need for vulnerability assessments to precede adaptation measures; the lack of locally Continued Danube: The large oods in the past decade appeared to have helped key institutions in the basin agree on the need for oodplain restoration and a key measure to manage the impacts of more frequent ooding due to climate change; Qualifying examples Godavari: The case study interventions appear sustainable after just 2 years. Yet this involved relatively local-scale actions applying one technology. It is likely that an iterative approach would be required to incrementally improve local benets, such as through tank watershed conservation, or to scale up tank restoration through larger programmes at the state or national scales.

CLIMATE AND DEVELOPMENT

Lessons for climate change adaptation from better management of rivers 203

TABLE 3 Continued Lesson derived Supporting examples available experts to advise on desirable measures; and the urgency of reducing existing threats to sustainable water management before engaging the longer-term impacts of climate change. 8. Post-disaster reform windows. Adaptation only followed major disasters or severe environmental degradation. All six projects demonstrate this (see Table 1). Qualifying examples Rio Conchos: Project leaders appear to have considered that the 19942006 drought may represent the sort of climate change impact to which their society would need to adapt. -

contrast, in the Yangtze and Danube, the oodwater retention capacities achieved by the restoration of oodplain sites are known and appear to be part of larger governmental decisions on the levels of acceptable ood risk.

4.2. Thresholds of climate impacts


Another key question is whether the resilience building measures implemented in these projects would be overwhelmed as climate change impacts exceed key thresholds. For instance, if climate change impacts become much more severe there is a risk that the responses to manage water scarcity and quality documented in these projects, successful to date, could be insufcient to meet the future water needs of people and the environment. Yet these resilience-building measures have engaged and enhanced the capacities of local institutions in adaptive management processes (such as with the Water User Associations and the Community Conservation Banks in the Great Ruaha) that may provide the social and institutional resources needed to respond to greater climate impacts. These actions to date have bought time to consider whether more radical measures are required. A recent assessment of the Ruaha basin was undertaken and has concluded that the improved catchment management measures (while highly benecial in increasing river ows and reducing vulnerability of local communities to water scarcity in the upper basin)

would not exceed the threshold needed to provide water ow through a major wetland and further downstream in the central basin in the dry season. Consequently the diversion of an upstream tributary, the Ndembera River, around the Usangu wetland has been proposed to provide a base ow to the main stem of the river (Mwaruvanda et al., 2009). Yet nearly all of the actions described in these case studies have two prized qualities: they are no or low regrets measures, for instance, in increasing ood retention capacity, and they can be scaled up considerably to substantially increase resilience at a basin or even greater scales, such as by restoring more of the estimated 208,000 village tanks in India. These case studies also illustrate the need to seize the impetus for adaptation following major disasters or severe environmental degradation (Adger et al., 2005, p. 85).

4.3. Motivation and adaptive behaviours


While WWF and local institutions did not initially conceive any of these projects as comprehensive climate change adaptation, the degree to which adaptation was considered varied between the projects. For example, the Danube project explicitly addressed oods as a climate change impact, while the Rio Sao Joao project had not thought of climate change until it was raised by this study. WWF is committed to promoting climate change adaptation measures globally

CLIMATE AND DEVELOPMENT

204 Pittock

and has provided guidance to staff since at least 2003 (Hansen et al., 2003). The project staff of WWF and their local institutional partners include people highly educated in relevant elds, yet few had focused on climate change. Why had these local intellectual leaders not fully considered adaptation needs? What would mobilize more leaders and their societies to mainstream adaptation processes? A response given by staff from three of the projects when approached to participate in this research was that insufcient climate change impact data was available for their river basin to enable development of targeted actions. Upon further discussion it appeared that there were two main reasons why planned climate adaptation work had not commenced in four of the six cases. There was a perception that measures could not commence until climate change models could supply more specic data on possible impacts, and a view that the required expertise and data was not available locally or nationally. In two cases project staff expressed the view that, because climate change projections included a range of contrasting potential outcomes in terms of rainfall and river inows, interventions could be premature. In general, the option of identifying no and low regrets adaptation measures had not been explicitly considered to manage risk and uncertainties. There was also a commonly held view (with some justication) that the current, non-climate threats to the sustainability of the river systems, such as excessive water diversions and pollution, were so large and fast-growing that they needed to be addressed rst. Thus most of the project staff saw the climate change information available to them as lacking salience (Meinke et al., 2006). Consequently, proponents of adaptation need to consider whether the sometimes obscure methods, jargon, data and expertise around climate adaptation are a barrier for many societies to implementation of appropriate measures; whether shifting from an information decit to more participatory approaches would be more effective in changing behaviour (Burgess et al., 1998). As a result of participating in this research,

the project staff responded with renewed condence that the adaptation actions they are implementing can be enhanced and become better climate-informed. This suggests that there are many local institutions that if directly engaged in locally relevant ways (Burgess et al., 1998; Meinke et al., 2006) will consider climate change adaptation measures. It was also clear that many local people and institutions initially implemented these actions more for the short-term benets for livelihood and development, and only later came to support the programmes for their benets in reducing vulnerabilities to climatic variability and other environmental hazards. To succeed, proponents need to link climate adaptation to outcomes of value to local communities.

4.4. Enhancing livelihoods vs. reducing risks


In terms of whether adaptation outcomes are best achieved by focusing on social and biophysical risk reduction or by development to reduce poverty and enhance livelihoods (Adger, 2006; Schipper, 2007) compelling evidence emerges from these projects that a concurrent investment facilitates more effective change. All the projects had substantial components focused on enhancing livelihoods as well as other environmental measures, consistent with WWFs support for sustainable development. This is well illustrated by the Great Ruaha and Godavari case studies, where average incomes were just USD0.80 and USD1.34 per day, respectively. In the Great Ruaha, the establishment of Community Conservation Banks enabled diversication into less water-intensive and more protable livelihoods concurrently with the establishment of Water User Associations to enhance local governance of water and reduce risks. In the Godavari basin, capital investment to reduce vulnerability to water scarcity by expanding village tanks was provided alongside the establishment of village tank management committees, the recruitment of local labourers to maintain the tanks, and enhancement of agricultural production and

CLIMATE AND DEVELOPMENT

Lessons for climate change adaptation from better management of rivers 205

employment. The prospect of sustainable development also appears to have been vital in securing the support of local and provincial scales of government, notably in China.

4.5. Centralized infrastructure vs. decentralized interventions


These case studies also highlight the benets of focusing on small-scale, decentralized adaptation measures and increasing societal capacities rather than investing in more technical infrastructure (Moench and Stapleton, 2007). In water management globally there is a pervasive bias towards investment in supply-side and centralized water infrastructure solutions (Molle, 2008). In the Godavari case, the WWF project suggests that restoring all the village tanks in the relatively small Maner River sub-basin at an estimated cost of USD635 million would provide a water storage capacity (1.764 billion m3 at 3 m depth) (Gujja et al., 2009) similar to the proposed USD4 billion Polavaram Dam on the Godavari River (2.130 billion m3) (Gujja et al., 2006). While not strictly comparable, because the larger catchment of the dam may see it divert over 4 times more water (8.130 billion m3 proposed) than the restored tanks, the cost of tank restoration appears cheaper than the proposed dam. Further, the dam would displace 250,000 people, inundate sites of environmental and cultural value, and supply a relatively constrained region. By contrast with the dam, restored village tanks would be more widely distributed to supply the poorest sections of society and enable local communities to manage their own water supplies. Similarly, in the cases of both the Danube and the Yangtze, the physical limits and costs of raising ever higher ood protection dykes appear to be outweighed by the potential of greater safety, lower costs and multiple benets derived from restoring oodplains.

4.6. Mainstreaming adaptation processes


Better methods for integrating adaptation processes into society are also suggested from the

WWF case studies which support Ross and Dovers (2008) proposition that the most prominent success factors, barriers and gaps that effect environmental policy integration relate to leadership, long term embedding of environmental policy integration and implementation capacity. Charismatic local leadership and the establishment or strengthening of local or basinscale institutions appears vital from the six projects. Local institutions had a key role in establishing new social norms to effect the changes in behaviour needed to better manage water, especially in those societies where the reach of government is limited, as illustrated in the Tanzanian, Indian and Brazilian case studies. At the basin scale, in the most successful programmes those in Brazil, Mexico and the Danube multi-stakeholder river basin organizations had been established and showed signs of the systematic social learning promoted by Lee (2003). Their work plans had set mediumterm targets that, once achieved, had built condence and facilitated a virtuous cycle of adaptive management through new iterations of basin plans similar to that described by Dovers (2005). In the case of Sao Joao, for instance, the Sao Joao Consorcio is implementing the third phase of their work, having achieved the pollution reduction and other targets of their rst two plans. In the best cases, a modest level of independent income appeared vital to the effectiveness of local institutions. In Brazil, the Sao Joao Consorcio secured a reliable income to underwrite a large part of their work from: (a) municipal government membership fees scaled according to the population of their jurisdictions, and (b) local company participation fees. In addition, the state government seconded a staff member to lead the secretariat of the consortium. By contrast, in Tanzania promises by the national government to allocate water user fees to local institutions had not been implemented, leaving these organizations dependent on aid funds. In Mexico, state law prevented the establishment of a

CLIMATE AND DEVELOPMENT

206 Pittock

scheme to provide compulsory payments for environmental services. The social, institutional and environment-focused interventions studied had a modest cost and were cheaper than either identied impacts or alternative options, such as the dam proposed in India. Upfront investment was required for necessary infrastructure, seed capital or loans, and to pay transition costs. The initial funding came from nongovernmental organizations, development banks and other aid donors. National governments often contributed funding only after the measures had shown the potential to succeed and, in some situations, were yet to implement promised reforms to guarantee funding for subnational institutions. This research sought to collect data on the return on investment of the measures implemented and proposed from these projects. While this information is incomplete, an assessment is possible. In the lower Danube a ood in 2005 killed 34 people and caused EUR396 million (USD625 million) in damages, whereas restoration of a larger area of oodplains would cost an estimated EUR20 million based on WWF project experience and generate ecological services worth EUR50 million per year. In Tanzania, each of the Water Users Associations cost USD13 27,000 to establish, and each Community Conservation Bank required an initial loan of USD4,000, compared to a national water budget of USD951 million from 2008 to 2011. The Maner sub-basin project to restore 12 village tanks serving 42,000 people cost WWF USD28,000 with USD75,000 contributed in kind by local people, and as indicated earlier, tank restoration appears more costeffective than the large dam proposed on the Godavari River. In the Rio Conchos, an initial expenditure of USD140 million was made to reduce irrigation surface water demand, and the Inter-institutional Working Group is investing USD3.2 4.4 million per year to manage the basin more sustainably. These examples demonstrate that investments in adaptations that reinforced institutional capacities and strengthened environmental resilience can be modest, may have a substantial return on investment,

and may be cheaper than alternative large-scale infrastructure projects.

4.7. Adaptation at different scales


Adger et al. (2005) argue that the success of adaptation processes at different scales can be judged by whether they are: (a) effective (are robust in the face of uncertainty and exible), (b) efcient (in terms of costs and benets), (c) equitable, and (d) legitimate. The most effective processes seen in these case studies drew strength and linked action at different geopolitical scales. Subnational governments were enthusiastic partners in these programmes, apparently motivated by sustainable development opportunities and the need to reduce vulnerability to natural hazards. National laws and resource provision that support basin- and sub-basin-scale institutions appeared vital for adaptive management of freshwaters, and were a considerable barrier where they did not exist, or had perverse impacts. Basin and multilateral treaties were a catalyst for better river management in trans-boundary situations, although they could be considered inexible in terms of their provisions in the event of climate change, for instance, in specifying particular water allocations. In the case of the Rio Conchos, Mexicos challenge in meeting its water delivery obligations to the USA ensured that considerable funding was available for river management reforms, that the state and national governments were supportive, and in future it may be possible to enhance environmental outcomes through smarter water delivery to the Rio Grande/Bravo. In the Danube basin the International Convention for the Protection of the Danube River, and the obligations of EU member states to implement the EU Water Framework Directive and related laws appeared to be a powerful driver for national law reform in Bulgaria and Romania. At different scales the measures adopted in the Danube are robust and exible in terms of, for instance, capacity to increase ood water retention. In each of these case studies the interventions appeared efcient,

CLIMATE AND DEVELOPMENT

Lessons for climate change adaptation from better management of rivers 207

with benets outweighing alternatives, social equity was improved through enhanced livelihoods, and legitimacy was established through extensive local and multi-stakeholder participation.

4.8. Responses from other practitioners


The Yangtze and Ruaha case studies and the preliminary conclusions of this research were presented on 21 August 2008 at the Water and Climate Day 2, Adaptation in Practice session of the Stockholm World Water Week, along with the research of four others. Dr Guy Howard, for the UK Department for International Development (Anon., 2008) summarized the main messages related to this research for more effective adaptation from the presentations and participants discussion. The importance of climatesmart local, regional and national water management institutions was recognized. Multiple benets were identied from investment in ecosystems as adaptation measures because they can be cheap, scalable and will not limit future options. It was noted that successful examples of self-help strategies relied on visible and relatively immediate socio-economic returns. The benets of learning by doing based on best current knowledge were observed, and the difculty of upscaling and mainstreaming adaptation strategies were noted. It was concluded that funding predictability is just as important as the scale of funding for local institutions.

5. Conclusions
The six cases studied in this report show that, when adaptation measures are considered in the context of common problems in water management, many practical ways of building resilience to climate change through mainstream programmes are evident. Many freshwater interventions identied in these projects could be scaled up and had benets for peoples livelihoods and for nature conservation: they were no and low regrets measures. Further lessons on

sub-programme detail and policy styles derived from these case studies that could support successful adaptation programmes include the value of: local ownership; provision of some immediate benets; linking local to global actions; applying adaptive management; consistently funding programmes; better explaining the opportunities for action; and seizing post-disaster policy response opportunities. This research demonstrates that adaptation is best considered as a pathway that starts by implementing no and low regrets measures to address obvious vulnerabilities that most societies could undertake with locally available knowledge and technologies. These small-scale measures can be scaled up, and they also buy time for thinking about and gathering the resources needed for solutions to more challenging problems, should they later emerge. A number of these case studies exhibit a virtuous cycle where initial, successful interventions have generated stakeholder support and built capacities for progressively more sophisticated measures that will further enhance adaptation to climate change. This pathway could be accelerated in many societies by investing in the development of expertise in technical skills and facilitating institutional development. Freshwater resources and ecosystems are under great threat from non-climate-related pressures, and water managers are focused on nding solutions to these challenges. The daunting and global nature of climate change appears to have further dissuaded many leaders and institutions from engaging in climate change adaptations in many of the cases studied. A common perception that particular expertise, data and methods are needed appears to have stalled active consideration of the issue and opportunities. This research supports Dovers (2009) propositions that we can go at least halfway to a believable adaptation policy by implementing known, well-supported policy and management options and by normalizing adaptation, and empowering ofcials, agencies, local communities.

CLIMATE AND DEVELOPMENT

208 Pittock

Acknowledgements
Dr Kossa Rajabu from WWF Tanzania, who died in December 2008, contributed to this research. WWF staff and local consultants who prepared the case studies assessed in this paper include:
B B B B

References
Adger, W. N., 2006. Vulnerability. Global Environmental Change, 16(3). 268 281. Adger, W. N., Arnell, N. W. and Tompkins, E., 2005. Successful adaptation to climate change across scales. Global Environmental Change Part A, 15(2). 77 86. Anon., 2007. The Brisbane Declaration. 10th International River Symposium and International Environmental Flows Conference. The Nature Conservancy, Brisbane, Australia. Anon., 2008. 2008 Synthesis Report. Progress and Prospects on Water: For a Clean and Healthy World with Special Focus on Sanitation. World Water Week, 17 23 August 2008. Stockholm International Water Institute, Stockholm. Barrios, J. E., Rodrguez-Pineda, J. A. and De la Maza, 2009. Integrated river basin management in the Conchos River basin, Mexico: a case study of freshwater climate change adaptation. Climate and Development, 1(3). 249 260. doi:10.3763/cdev.2009.0024. Bates, B. C., Kundzewicz, Z. W., Wu, S. and Palutikof, J. P. (eds), 2008. Climate Change and Water. Technical Paper of the Intergovernmental Panel on Climate Change. IPCC Secretariat, Geneva. Brooks, N., 2003. Vulnerability, Risk and Adaption: A Conceptual Framework. Tyndall Centre for Climate Change Research Working Paper 38. Tyndall Centre for Climate Change Research, Norwich, UK. Burgess, J., Harrison, C. M. and Filius, P., 1998. Environmental communication and the cultural politics of environmental citizenship. Environment and Planning A, 30. 1445 1460. Dovers, S., 2005. Environment and Sustainability Policy: Creation, Implementation, Evaluation. Federation Press, Annandale, VA. Dovers, S., 2009. Normalizing adaptation. Global Environmental Change, 19(1). 4 6. Dryzek, J. S., 1997. The Politics of the Earth: Environmental Discourses. Oxford University Press, Oxford and New York. Ebert, S., Hulea, O. and Strobel, D., 2009. Floodplain restoration along the lower Danube: a climate change adaptation case study. Climate and Development, 1(3). 212219. doi:10.3763/cdev.2009.0022. Gujja, B., Ramakrishna, S., Goud, V. and Sivaramakrishna (eds), 2006. Perspectives on Polavaram: A Major Irrigation Project on the Godavari. Academic Foundation, New Delhi. Gujja, B., Dalai, S., Shaik, H. and Goud, V., 2009. Adapting to climate change in the Godavari River basin of India by restoring traditional water storage systems.

B B

Danube: Andreas Beckmann, David Strobel, Suzanne Ebert and Kimberly Chan. Ruaha: Japhet Kashaigili, Petro Masolwa and Kossa Rajabu. Godavari: Biksham Gujja, Sraban Dalai, Hajara Shaik and Vinod Goud. Yangtze: Xiubo Yu, Luguang Jiang, Jinxin Wang, Jiang Zhu, Gang Lei, Limin Wang and Lifeng Li. Sao Joao: Firmino Pereira, Samuel Barreto and Michael Volcker. Conchos: Venancio Trueba, Eugenio Barrios, Mauricio De la Maza Benignos and Alfredo Rodrguez.

This research was sponsored by HSBC Climate Partnership, and was supported by WWF UK, in particular: Dave Tickner, Philip Leonard, Tom Le Quesne and Mica Ruiz. Comments were gratefully received from Dr John Matthews, Dr Karen Hussey and Prof Stephen Dovers, and two reviewers. This research draws on the work of many WWF staff, partners and donors whose contributions are greatly appreciated.

Notes
1. Resilience has been dened as the ability of a social or ecological system to absorb disturbances while retaining the same basic structure and ways of functioning, the capacity for self organization, and the capacity to adapt to stress and change (IPCC, 2007). 2. Vulnerability has been dened as the degree to which a system is susceptible to, and unable to cope with, adverse effects of climate change, including climate variability and extremes. Vulnerability is a function of the character, magnitude, and rate of climate change and variation to which a system is exposed, its sensitivity, and its adaptive capacity (IPCC, 2001).

CLIMATE AND DEVELOPMENT

Lessons for climate change adaptation from better management of rivers 209

Climate and Development, 1(3). 229240. doi:10.3763/ cdev.2009.0020. Hansen, L. J., Hoffman, J. R. and Biringer, J. L., 2003. Buying Time: A Users Manual to Building Resistance and Resilience to Climate Change in Natural Systems. WWF, Gland, Switzerland. Head, B. W., 2008. Three lenses of evidence-based policy. Australian Journal of Public Administration, 67(1). 111. IPCC (Intergovernmental Panel on Climate Change), 2001. Climate Change 2001: Impacts, Adaptation, and Vulnerability: Contribution of Working Group II to the IPCC Third Assessment Report 2001, J. J. McCarthy, O. F. Canziani, N. A. Leary, D. J. Dokken and K. S. White (eds). Cambridge University Press, Cambridge, UK. IPCC (Intergovernmental Panel on Climate Change), 2007. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Appendix I: Glossary, M. L. Parry, O. F. Canziani, J. P. Palutikof, P. J. van der Linden and C. E. Hanson (eds). Cambridge University Press, Cambridge, UK. 869 883. Kashaigili, J. J., Rajabu, K. and Masolwa, P., 2009. Freshwater management and climate change adaptation: experiences from the Great Ruaha River catchment in Tanzania. Climate and Development, 1(3). 220228. doi:10.3763/cdev.2009.0025. Kundzewicz, Z. W., Mata, L. J., Arnell, N. W., Doll, P., Kabat, P., Jimenez, B., Miller, K. A., Oki, T., Sen, Z. and Shiklomanov, I. A., 2007. Freshwater resources and their management. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, M. L. Parry, O. F. Canziani, J. P. Palutikof, P. J. van der Linden and and C. E. Hanson (eds). Cambridge University Press, Cambridge, UK. 173210. Lee, K. N., 1993. Compass and Gyroscope: Integrating Science and Politics for the Environment. Island Press, Washington, DC and Covelo, CA. Matthews, J. H. and Wickel, A. J., 2009. Embracing uncertainty in freshwater climate change adaptation: a natural history approach. Climate and Development, 1(3). 269279. doi:10.3763/cdev. 2009.0018. Meinke, H., Nelson, R., Kokic, P., Stone, R., Selvaraju, R. and Baethgen, W., 2006. Actionable, climate knowledge: from analysis to synthesis. Climate Research, 33. 101 110. Moench, M. and Stapleton, S., 2007. Water, Climate, Risk and Adaptation, Working Paper 2007/01. Cooperative Programme on Water and Climate, Delft, The Netherlands.

Molle, F., 2008. Why enough is never enough: the societal determinants of river basin closure. Water Resources Development, 24(2). 217 226. Mwaruvanda, W., Kashaigili, J. J., OKeeffe, J. H. and Von Der Heyden, C., 2009. Restoring ows in the Great Ruaha River, Tanzania. International Conference on Implementing Environmental Water Allocations. Port Elizabeth, South Africa. Pereira, L. F. M., Barreto, S. and Pittock, J., 2009. Partici patory river basin management in the Sao Joao River, Brazil: a basis for climate change adaptation? Climate and Development, 1. 18. doi:10.3763/cdev.2009.026. Ribot, J. C., Najam, A. and Watson, G., 2009. Climate variation, vulnerability and sustainable development in the semi-arid tropics. The Earthscan Reader on Adaptation to Climate Change, E. L. F. Schipper and I. Burton (eds). Earthscan, London. 117 160. Ross, A. and Dovers, S., 2008. Making the harder yards: environmental policy integration in Australia. Australian Journal of Public Administration, 67(3). 245 260. Schipper, E. L. F., 2007. Climate Change Adaptation and Development: Exploring the Linkages. Tyndall Centre Working Paper No. 107. Tyndall Centre for Climate Change Research, Norwich, UK. Yu, X., Jiang, L., Li, L., Wang, J., Wang, L., Lei, G. and Pittock, J., 2009. Freshwater management and climate change adaptation: experiences from the central Yangtze in China. Climate and Development, 1(3). 241 248. doi:10.3763/cdev.2009.0023.

Annex: Analytical framework A. Background and overview


Place/river basin Country Why it is an example of climate change adaptation Summary (of sections B D):
B B B

Change in climate change resilience Change in livelihoods Change in conservation status

Key lessons
B B

What worked What did not work well

CLIMATE AND DEVELOPMENT

210 Pittock

Timeline of processes and WWF and partners interventions Quality of the data. If any of the questions below could not be answered, why not? Main actors their roles and relationships: Government agencies: local/provincial or state/national/multilateral B Business B Community B Multi-stakeholder
B

What intra- and inter-governmental processes were used? What elements made interactions between these stakeholders positive or negative?

B. Climate change adaptation


1. What was the baseline situation? 2. What are the natural historic, climatic and hydrologic risks in the area? 3. How do local people cope with these risks traditionally? 4. What increased risks are forecast with climate change? 5. What are the projects climate change adaptation outcomes? Can these be quantied? To what extent are these based on having more resilient institutions? 6. Were these planned or serendipitous? 7. Were these planned to address a future forecast threat (e.g. potentially larger oods or greater water scarcity) or were they intended to incrementally improve management of an existing problem (e.g. current ood levels or current water shortages)? 8. Is the improvement in climate change adaptation sustainable? 9. Why has the project been successful in improving climate change adaptation? B What activities have been carried out at the macro, meso and micro scales? B Which formal institutions were in place that have contributed to a favourable outcome?

Which informal institutions were in place that have contributed to a favourable outcome? B What assumptions were made before the project was implemented and were these realistic? B What was the time frame within which benets could be measured? B How were local people, their knowledge and needs integrated into the project? B Which partnerships with stakeholders were established in the project and what roles did these play? 10. What should be done differently for similar projects in future? 11. Can the outcomes in the project site compare to a similar place that was not involved in the project? 12. What needs to be done to ramp up these adaptation techniques to the basin scale and what would it cost?
B

C. Socio-economics
1. What was the baseline situation? 2. What are the projects livelihood outcomes? (a) More income? (b) Increased well-being? (c) Reduced vulnerability? (d) More sustainable use of the natural resource base? 3. What is the distribution of socio-economic benets? (a) Gender? (b) Age groups? (c) Income groups? (d) Disadvantaged groups (HIV/AIDS, unemployed, disabled, etc)? 4. What would have happened to peoples livelihoods without the project? 5. Is the improvement in livelihoods sustainable? 6. Why has the project been successful in improving livelihoods? B What activities have been carried out at the macro, meso and micro scales?

CLIMATE AND DEVELOPMENT

Lessons for climate change adaptation from better management of rivers 211

Which formal institutions were in place that have contributed to a favourable outcome? B Which informal institutions were in place that have contributed to a favourable outcome? B What assumptions were made before the project was implemented and were these realistic? B What was the time frame within which benets could be measured? B How were local people, their knowledge and needs integrated into the project? B Which partnerships with stakeholders were established in the project and what roles did these play? 7. What should be done differently for similar projects in future? 8. Can you compare the outcomes in the project site to a similar place that was not involved in the project? 9. What would be the socio-economic impacts of business-as-usual and what is the benet of magnifying the project to the other relevant parts of the river/basin? Balance this with what it would cost to implement these adaptation techniques in the above section.
B

D. Conservation
1. What was the baseline situation? 2. What was the conservation objective/s of WWFs intervention/s?

3. What are the projects environmental outcomes? 4. Is the improvement in conservation sustainable? 5. Why has the project been successful in improving conservation? B What activities have been carried out at the macro, meso and micro scales? B Which formal institutions were in place that have contributed to a favourable outcome? B Which informal institutions were in place that have contributed to a favourable outcome? B What assumptions were made before the project was implemented and were these realistic? B What was the time frame within which benets could be measured? B How were local people, their knowledge and needs integrated into the project? B Which partnerships with stakeholders were established in the project and what roles did these play? 6. What should be done differently for similar projects in future? 7. Can you compare the outcomes in the project site to a similar place that was not involved in the project? 8. What would be the impact of business-asusual and the conservation/ecological benets of ramping up to the river/basin scale?

CLIMATE AND DEVELOPMENT

case study

Floodplain restoration along the lower Danube: A climate change adaptation case study
SUZANNE EBERT1,*, ORIETA HULEA2 and DAVID STROBEL1
1 2

WWF Danube-Carpathian Programme Ofce, Mariahilferstrasse 88a/3/9, 1070 Vienna, Austria WWF Danube-Carpathian Programme Romania, Mircea Vulcanescu Street No. 109 Sector 1, Bucharest, RO-01 0818 Romania

Conversion of the Danube river oodplains through dyke construction for farming and other development has cut off 95, 75 and 28% of the oodplains of the upper Danube, the lower Danube and the Danube delta, respectively. Together with channelization, this has exacerbated ood peaks. Anthropogenic climate change is anticipated to bring more frequent ooding and reduced water quality. In assessing ongoing oodplain restoration work that commenced in 1993, this paper nds the following. (a) Along the lower Danube River, restoration of oodplains by decommissioning under-performing ood protection infrastructure has provided many benets. The benets of these adaptation measures include improved natural capacity to retain and release oodwaters and remove pollutants, enhanced biodiversity, and strengthened local economies through diversication of livelihoods based on natural resources. (b) The drivers for more successful adaptation measures in the Danube included EU expansion, legal mechanisms, and local desire to improve livelihoods. The support of non-governmental organizations (WWF and partner organizations) for basin- and regional-level planning for more effective water resource management has also been a powerful driver of policy change in the lower Danube countries. Keywords: climate change adaptation; Danube; oodplain restoration; oods; Romania; Ukraine

1. Introduction
This paper assesses freshwater climate change adaptation work in the lower Danube River basin in Romania and Ukraine, instigated by WWF, a conservation non-governmental organization (NGO). The purpose of this assessment is to derive lessons on (a) what motivated policymakers to act, (b) which factors led to more successful climate adaptation, and (c) how the interventions may be best sustained in coming decades. We consider that better practices for adaptation may best be identied from existing adaptation measures. Hence this study was undertaken to draw lessons from work that commenced in 1992 to restore oodplains and reduce climate change impacts on freshwater ecosystems and the livelihoods dependent on these environments.

Falling within the territories of 19 European states, the 801,463 km2 Danube River basin is home to 81 million people (ICPDR, 2004). At 2,780 km long, the Danube River (see Figure 1) has been subject to extensive development and political change. Based on the gradients of different sections, the Danube River can be divided into three sub-regions: the upper basin from the source to Bratislava in Slovakia, the middle basin from Bratislava to Iron Gates in Romania, and the lower basin from Iron Gates to the Danube Delta on the Black Sea (see Figure 2). This paper focuses on the lower Danube basin. This paper also assesses whether restoration of natural oodplain integrity a soft adaptation in place of conventional hard infrastructure solutions provides greater benets to nature and human livelihoods and more

B *Corresponding author. E-mail: sebert@wwfdcp.org


CLIMATE AND DEVELOPMENT 1 (2009) 212219 doi:10.3763/cdev.2009.0022 # 2009 Earthscan ISSN: 1756-5529 (print), 1756-5537 (online) www.earthscanjournals.com

Floodplain restoration along the lower Danube 213

FIGURE 1 The Danube River basin

long-term exibility in addressing negative impacts from anthropogenic climate change. Resilient, healthy habitats such as wetlands and natural river side arms not only aid biodiversity conservation, but also enhance the services these ecosystems supply to local people, such as better water quality, sh, reeds and timber. Pollution from agriculture (50%), cities (25%) and industry (25%) makes the Danube

the largest source of nutrients into the Black Sea, which suffers from a hypoxic dead zone near the estuary (Behrendt, 2008). Restoring riverine wetlands may be one means of reducing this pollution. Conversion of historical oodplains through ood protection dykes for agriculture, aquaculture and intensive forestry has cut off 95, 75 and 28% of the oodplains of the upper Danube, the

CLIMATE AND DEVELOPMENT

214 Ebert, Hulea and Strobel

FIGURE 2 Boundaries of the upper, middle and lower Danube River basin

lower Danube and the Danube delta, respectively (UNDP/GEF, 1999). The gradient of the remaining oodplain reects actions over the course of the 20th century. In particular, the lesser degree of development along the lower Danube in the communist era left intact large areas of oodplain forest and other wetland habitats. The excision of the oodplains, especially upstream, has exacerbated ood peaks. In 2005 a ood killed 20 people, displaced 10,000 people, and caused USD625 million (EUR444 million) in damages in Bulgaria (Petrova, 2005). One year later a ood displaced 15,000 people and inundated 80,000 ha in Romania alone (Shepherd, 2005). Moreover, oods in Romania have caused an estimated EUR1.66 billion in damages between 1992 and 2005, exceeding the gross national product (GNP) by 0.6% (Mihailovici, 2006). Climate change is expected to lead to major changes in annual and seasonal water availability across

Europe. South-eastern regions (including the lower Danube region) will be particularly exposed to an increase in the frequency and intensity of droughts as well as extreme high river ows due to an increase in heavy rain events (Czako and Mnatsakanian, 2008; EEA, 2008). Projections for Romania (Balteanu et al., 2009) show an expected increase in mean annual temperature of 2C over the next 40 years, as well as signicant seasonal variability of the precipitation regime. Although quantitative projections of changes in precipitation and river ows remain uncertain (EEA, 2008), climate change signals are sufcient to justify action beyond existing scientic uncertainties. Along the lower Danube River more frequent ooding is expected. Reactivation of former wetlands and oodplain may increase oodwater retention and improve water quality, thus beneting nature and the people of the region.

CLIMATE AND DEVELOPMENT

Floodplain restoration along the lower Danube 215

2. Methods
WWF commenced work on the Danube in 1992 and promoted the establishment of the Convention for the Protection of the Danube River (DRPC, 1994) in 1994 and the European Union (EU) Water Framework Directive (EC, 2000) in 2000. Also in 2000, WWF secured agreement from the heads of state of Bulgaria, Romania, Moldova and Ukraine to restore 2,236 km2 of oodplain to form the 9,000 km2 Lower Danube Green Corridor (or LDGC) (WWF, 2008). The LDGC is intended to attenuate oods, restore and protect biodiversity, improve water quality and enhance local livelihoods. The projects investigated here are located in the LDGC area. Pilot projects to demonstrate the importance of oodplain restoration assessed in this case study include: the 1993 1996 reconnection of the 36.8 km2 Babina and Cernovca polders to the Danube hydrological regime (i.e. a low-lying tract of land used for agriculture or sh farming, usually separated from a nearby body of water by embankments) in Romania, and B the re-linking of the 68 km2 Katlabuh Lake to the river and removal of dykes on the 7.5 km2 Tataru Island in Ukraine from 2005 to 2008.
B

offers natural spawning and nesting grounds for sh and birds among other important natural resources for the local communities. Although accounting for a small part of the LDGC potential restoration area, these pilot projects demonstrate the benets of restoring lost and degraded wetlands and provide valuable experience and lessons for further restoration initiatives. In 2006, the potential for oodplain restoration and the potential costs and benets of using soft infrastructure for ood protection along the Danube were assessed by WWF (Schwarz et al., 2006). The study used GIS data and satellite images to evaluate oodplain loss along the Danube. Detailed analyses of four smaller areas on the lower Danube were used to better quantify the costs and benets of oodplain restoration on ood risk mitigation, including the retention area and discharge capacity. This 2009 case study was prepared using an analytical framework as part of a larger review of freshwater autonomous adaptation projects by WWF (see Pittock, 2009).

3. Results
As of 2008, 469 km2 of oodplain 14.4% of the LDGC area pledged by the governments in 2000 has been restored or is undergoing restoration (WWF Danube-Carpathian Ofce, unpublished data). Although implementation of the full oodplain restoration is incomplete, ood control benets are already visible. The restored 21 km2 Babina island polder holds 35 million m3 in oodwaters during signicant inundation events (Marin and Schneider, 1997). These pilot restoration sites are in the Danube delta where the ood safety benets are less obvious than at sites located further upstream, but they clearly demonstrate the value of restoring oodplains to lessen the impacts of ood events. If the 2000 LDGC agreement to restore a total area of 2,236 km2 is fully implemented, potential ood control benets would be even larger. Moreover, the restoration of oodplains and former side channels along the entire Danube, not just in the LDGC area, would provide nearly 2,100 million m3

Restoration of Babina and Cernovca polders was possible due to cooperation between WWF and the Danube Delta Research Institute. The opening of the surrounding dykes and natural ooding of the polders resulted in a mosaic of aquatic habitats rich in biodiversity, natural resources and water retention areas, which in turn provided benets to the local communities. In Ukraine, as part of a partnership between WWF and the Odessa Oblast Water Management Board, a sluice was built in the dyke separating Katlabuh Lake from the Danube River. WWF and the local forestry unit joined efforts to remove the surrounding dykes on Tataru Island, allowing natural ooding to occur to revitalize the wetlands and the oodplain forest. The island, formerly a forestry polder uneconomic as a business, now

CLIMATE AND DEVELOPMENT

216 Ebert, Hulea and Strobel

in ood retention capacity and would lower Danube extreme ood peaks (like the 2006 oods) by 40 cm (Schwarz et al., 2006, p. 5). To aid decision making by governments on oodplain restoration priorities, WWF has identied potential oodplain restoration sites throughout the Danube basin that coincide with biodiversity conservation priorities, whose restoration offers dual biodiversity conservation and ood control benets (Schwarz et al., 2006, p. 30). From a development perspective, oodplain restoration appears to enhance local livelihoods. Reduced vulnerability to oods by restoring the retention capacity of the oodplain, especially by reconnecting side arms and widening the oodplain upstream of settlements, is a major benet for communities. Most of the polders targeted for conversion within the Danube delta were used for intense cropping, an activity that was neither very appropriate for the local environmental conditions nor protable since the change from centralized economies in the 1990s. Most polders were associated with declining prots over recent decades due to poor land management (Staras, 2000). Based on data from Stiuca et al. (2002), restoration of the Babina and Cernovca pilot polders in Romania resulted in a diversication in livelihood strategies towards shing, tourism, reed harvesting and livestock grazing on seasonal pastures, activities that earn an average USD37 per ha per year (about USD140,000/year for both polders; EUR26 per ha/year and EUR99,000/year). From an ecosystem perspective, each hectare of restored wetland is calculated to produce 34 kg of commercial-sized sh per year (Stiuca et al., 2002), and at the 36.8 km2 Babina and Cernovca polders, the restored sheries provide jobs for 20 25 people (Staras, 2000). At Katlabuh Lake, improved water quality will enhance access for 10,000 local residents to drinking and irrigation water. Natural wetland habitats have returned to Tataru Island after dykes were removed. According to Kettunen and ten Brink (2006), a large-scale calculation of the economic values for the restored lower Danube estimate the benets based on Romanian expert estimations for

nutrient reduction, provision of sh, reeds, crops, vegetables, animals and tourism at EUR1,354 per ha/year. Schwarz et al. (2006) estimate economic benets from nutrient reduction in oodplains at EUR870 per ha/year. Another WWF study calculates the value based on provision of sh, forestry, animal fodder, nutrient retention as well as recreation and gives an estimate of about EUR383 per ha/year (Gren et al., 1995). Therefore, based on these highly differing economic values, an average value was calculated to be around EUR500 per ha/year (Schwarz et al., 2006) for provision of ecosystem services for sheries, forestry, animal fodder, nutrient retention and recreation through oodplain restoration. Limited data availability and published studies on the economic value of the current land use in the oodplain does not allow for effective comparison with the values of restored functions and services of the wetlands. However, recent assessments in Romania (DDNI, 2008) show an economic value of EUR360 per ha/year for areas intensively used for agriculture. If the total pledged oodplain area in the LDGC were restored, we estimate the value of the resulting additional ecosystem services at EUR111.8 million annually (225,000 ha EUR500/ha). Restoration efforts in the 9,000 km2 LDGC also enhanced biodiversity conservation. For instance, following restoration of the 21 km2 Babina island polder, the number of resident bird species increased from 34 to 72, and over a quarter of the water bird species commenced breeding (Marin and Schneider, 1997).

4. Discussion
Motivations for oodplain and wetland restoration varied among stakeholder groups. Local communities directly relying on the availability and quality of natural resources were supportive of restoration measures likely to improve their livelihood and bring opportunities to strengthen local economies. National governmental authorities undertook restoration to improve the local ecological

CLIMATE AND DEVELOPMENT

Floodplain restoration along the lower Danube 217

situation, reduce vulnerability to ooding, improve water quality and increase local incomes while reducing the pressures on the natural areas. NGOs, such as WWF, support restoration to promote biodiversity conservation through habitat improvement, among other things. National governments are seeking to full their obligations under EU legislation, such as the Water Framework Directive and the Danube River Protection Convention, to adopt new and more sustainable river management practices. The expansion of the EU into Eastern Europe has been one driver for the reform of river basin management, to promote integrated water resources management, together with the obligation of Eastern European countries (Romania and Bulgaria) to transpose and comply with EU laws. This reinforces the points Adger et al. (2005) make on the need for and benets of implementation of adaptation measures across scales. The oodplain restoration efforts appear to be sustainable. Reversion to previous exploitative attitudes towards oodplain habitats is unlikely in most cases because of the high cost of rebuilding dykes; in most cases the restored oodplains are designated as protected areas (of national or European importance), local peoples livelihoods have improved and the threat from ooding has lessened. Management costs of the restored oodplains are low relative to the hard infrastructure they replace. However, political barriers were encountered during the restoration process. Government implementation at the national and local levels of restoration efforts was slower than anticipated. The appointment of ofcials and agencies to lead the work is time-consuming, which exacerbated the time required to develop national implementation plans and allocate restoration funds. Most of the funding for oodplain restoration in the pilot areas has come from a combination of local authorities, EU, NGOs and other donor organizations. In some instances (especially in the LDGC area) local people do not always consent to restoration where changes in land ownership and concession laws have hindered

progress. To reduce these obstacles, WWF has informed and lobbied stakeholders at local, national and international levels for oodplain restoration, signing memorandums of understanding and organizing public meetings and seminars. The organization has also resorted to providing resources to cover costs ineligible for governmental funds. In one case WWF provided up to one-half of total restoration costs for co-funding a pilot project. Making use of post-disaster policy windows has been a key lesson. Governments of countries affected by the 2005 2006 oods took immediate actions to develop ood risk mitigation strategies and action plans that include oodplain restoration as an adaptation measure. For instance, Romania is currently completing a national oodplain restoration strategy aimed at reducing ood risk (DDNI, 2008). In spite of favourable cost benet analyses (UNDP/GEF, 1999), policymakers have been slow to promote oodplain restoration as a form of sustainable ood protection because: it is still a relatively new concept in the region and lacks some governmental commitment; B changing the land use in the area has socioeconomic implications (most of the land has been privatized); B inuential stakeholders with interests in short-term exploitation of the oodplain exert political pressure; and B the cost of transitional restoration works and compensatory measures is high.
B

Individual actions are constrained by institutional processes, such as regulatory structures, property rights, and social norms associated with rules in use (Adger et al., 2005). This is true for the examples presented in this case study, where the success of the interventions depends highly on the constraints on smaller political and jurisdictional scales, such as the municipality, county or concession area. Persistent work over more than 10 years has been required to achieve the outcomes to date. Considerable time and resources for monitoring

CLIMATE AND DEVELOPMENT

218 Ebert, Hulea and Strobel

and assessment in terms of adaptive management should be major components for other groups considering similar projects. Linking work at local, national, basin and European scales has been critical to achieving change. There is great potential to scale up restoration activities from these pilot projects. Based on the Romanian pilot projects WWF estimates that dyke removal costs EUR50,000 200,000 per km, depending on the nature of the dyke wall, plus compensation for changes in land use (Schwarz et al., 2006). From this work WWF has calculated that restoration of four polders covering 1,000 km2 in Romania would cost around EUR20 million, hold 1,600 million m3 of oodwaters, and generate ecosystem services worth EUR50 million per year (Schwarz et al., 2006). These ecosystem services do not include an estimate of the losses that would result from ooding if these sites were unrestored, and signicantly these four polders did ood in 2006. Further, restoration of the 37 oodplain sites that make up the LDGC is estimated to cost EUR183 million (WWF Danube-Carpathian Programme, unpublished data), compared to total damages costing an estimated EUR400 million on the lower Danube from the spring 2006 oods (Schwarz et al., 2006) and likely ecosystem services earnings of EUR111.8 million per year (see above). Clearly, oodplain restoration is a cost-effective approach that can be expanded across the Danube basin. Major factors affecting the implementation of restoration potential are weak political and governmental commitment and unbalanced allocation of nancial resources to infrastructure measures. The projects described here were only partly designed to address climate change impacts in the region. High condence climate change scenarios that would enable managers to adopt specic countermeasures are not available for the Danube River basin. An analysis of historical extremes and regional climate projections, such as those presented in the recent IPCC Technical Paper on Climate Change and Water (Bates et al., 2008), indicates that more frequent and severe oods are likely. Restoration of oodplains

serves as a no regrets form of climate change adaptation measure by attenuating the impacts of oods, aiding groundwater recharge that may maintain water supplies during droughts, and improving water quality through wetland ltration of pollutants.

5. Conclusions
This assessment of large-scale restoration of oodplain functions in the lower Danube basin, partly as a climate change adaptation measure, showed that the benets of using soft infrastructure for multiple purposes outweighed the costs. Decommissioning underperforming ood protection dykes and restoring oodplains led to safer and more effective oodwater retention, more robust and dependable freshwater ecosystem services, lower infrastructure maintenance costs and ultimately more sustainable development trajectories for these emerging economies. We contend that these no regrets measures increase resilience of natural systems and local societies in managing current climate variability and the likely impacts of further climate change. Successful restoration of agriculture and forestry polders have replaced vulnerable monocultures with more diverse and resilient livelihoods based on sustainable ecosystem services, directly beneting the tourism, shing, grazing and bre production industries. A number of lessons are derived from this case study that may guide adaptation priorities for governments and aid donors. The development of international institutions and agreements (e.g. International Commission for the Protection of Danube River ICPDR, EU Water Framework Directive) for better water and river management have been powerful drivers for more sustainable management of the Danube River. Finally, increasing awareness of the socio-economic benets of oodplain restoration, and the demonstrated inefciency of existing ood protection infrastructure in coping with extreme climate events, are key factors proven to motivate policymakers to start integrating climate change

CLIMATE AND DEVELOPMENT

Floodplain restoration along the lower Danube 219

into national and regional development strategies. We hope that these successes achieved to date will provide the basis for further adaptive management, particularly as the impacts from climate change grow in frequency and severity.

References
Adger, W. N., Arnell, N. W. and Tompkins, E., 2005. Successful adaptation to climate change across scales. Global Environmental Change A, 15(2). 77 86. Balteanu, D., Dumitrascu, M. and Micu, D., 2009. Impact of Climate Change on Wetlands. Presentation at the BIOWETMAN International Symposium, 19 February 2009, Bucharest, Hungary. Bates, B. C., Kundzewicz, Z. W., Wu, S. and Palutikof, J. P. (eds), 2008. Climate Change and Water. Technical Paper of the Intergovernmental Panel on Climate Change. IPCC Secretariat, Geneva. Behrendt, H., 2008. Nutrient Reduction Scenarios for the Danube River Basin District (DRBD) and First Assumptions for the Baseline Scenario for 2015. Institute of Freshwater Ecology and Inland Fisheries, Berlin. Czako, V. and Mnatsakanian, R., 2008. Impacts of and Adaptation to Climate Change in the Danube Carpathian Region: Overview Study Commissioned by the WWF DanubeCarpathian Programme. Central European University, Budapest, Hungary. DDNI, 2008. Ecological and economic rehabilitation of the Lower Danube oodplain in Romania. Research Report No. 7420.73: study commissioned by the Ministry of Environment and Sustainable Development (unpublished data). DRPC, 1994. Convention on Cooperation for the Protection and Sustainable Use of the Danube River (Danube River Protection Convention), 29 June 1994. International Convention for Protection of the Danube River, Soa, Bulgaria. EC, 2000. Directive 2000/60/EC of the European Parliament and of the Council of 23 October 2000 establishing a framework for Community action in the eld of water policy, nal. European Commission, Brussels. EEA, 2008. Impact of Europes Changing Climate: 2008 Indicator Assessment. European Environment Agency Report No. 4/2008. EEA, Copenhagen. Gren, I.-M., Groth, K.-H. and Sylven, M., 1995. Economic values of Danube oodplains. Journal of Environmental Management, 45(4). 333 345.

ICPDR, 2004. Danube Basin Analysis (WFD Roof Report 2004). International Convention for Protection of the Danube River Document IC/084 (version 1). ICPDR, Vienna. Kettunen, M. and ten Brink, P., 2006. Value of Biodiversity: Documenting EU Examples Where Biodiversity Loss has led to the Loss of Ecosystem Services. Final report for the European Commission. IEEP, Brussels. Marin, G. and Schneider, E., 1997. Ecological Restoration in the Danube Delta Biosphere Reserve, Romania: Babina and Cernovca Islands. ICPDD (Danube Delta Research and Design Institute, Romania). WWFAueninstitut, Rastatt, Germany. Mihailovici, M., 2006. Floods in Romania 2005 2006. Presentation at the European Flood Conference, 17 18 May 2006. Vienna. Petrova, K., 2005. Flood Crisis Hits Bulgaria Hard. BBC News Online, press article. http://news.bbc.co.uk/ 1/hi/world/europe/4180996.stm. Pittock, J., 2009. Lessons for climate change adaptation from better management of rivers. Climate and Development, 1(3). 194211. doi: 10.3763/cdev.2009.0021. Schwarz, U., Bratrich, C., Hulea, O., Moroz, S., Pumputyte, N., Rast, G., Bern, M. R. and Siposs, V., 2006. 2006 Floods in the Danube River Basin: Flood Risk Mitigation for People Living along the Danube and the Potential for Floodplain Protection and Restoration. Working paper. WWF Danube-Carpathian Programme, Vienna. Shepherd, K., 2005. Flooding Fears Return to the Danube. ICPDR Danube Watch, Issue 2/2006, press article. www.icpdr.org/icpdr-pages/dw0602_p_07.htm. Staras, M., 2000. Restoration programme in the Danube Delta: achievements, benets and constraints. Proceedings from the Conference on River Restoration. Wageningen, The Netherlands, . Stiuca, R., Staras, M. and Tudor, M., 2002. Ecological restoration in the Danube Delta: alternatives for sustainable management of degraded wetlands. Proceedings from Limnological Reports (Vol. 34) of the 34th Conference of the International Association for Danube Research (IAD), 2731 August 2002. Tulcea, Romania. UNDP/GEF, 1999. Evaluation of Wetlands and Floodplain Areas in the Danube River Basin. Danube Pollution Reduction Programme (DPRP), Vienna. WWF, 2008. Lower Danube Green Corridor. www. panda.org/what_we_do/where_we_work/black_sea_ basin/danube_carpathian/our_solutions/freshwater/ oodplains/lower_danube_and_danube_delta/.

CLIMATE AND DEVELOPMENT

case study

Freshwater management and climate change adaptation: Experiences from the Great Ruaha River catchment in Tanzania
JAPHET J. KASHAIGILI1, *, KOSSA RAJABU2 and PETRO MASOLWA2
1 2

Faculty of Forestry and Nature Conservation, Sokoine University of Agriculture, P.O. Box 3013, Morogoro, Tanzania Ruaha Water Programme, World Wide Fund for Nature Tanzania Programme Ofce, P.O. Box 307, Iringa, Tanzania

Adaptation to anthropogenic climate change is becoming vital to freshwater ecosystems and resource management, but climate adaptation can be purposeful or unintentional. This paper presents lessons from an assessment of an autonomous adaptation in the Great Ruaha River catchment in Tanzania following WWF intervention. The project was designed to address challenges resulting from natural resource use and existing levels of climate variability by changing water resource management. The study applied participatory methods and an open-ended questionnaire to collect data. The study found key adaptation benets, including reduced vulnerability to drought and strengthened local water user associations and other regional institutions. As a result of the project interventions from 2003 to 2009, rural livelihoods became more protable and water-sustainable, and local livelihood strategies were diversied. Regional ecosystems improved as a result of restoring river ows in some rivers, conservation of riparian vegetation and halting tree felling for charcoal production. As a result of these changes the communities and ecosystems in the Great Ruaha River catchment should be more resilient to emerging climate change impacts, yet the need for further physical interventions and institutional reform is identied. The study concludes that strengthening local institutions and capacity building are fundamental to climate change adaptation and sustainable freshwater management. Keywords: climate change adaptation; climate vulnerability; freshwater management; Great Ruaha River; institutions; sustainability

1. Introduction
Freshwater is essential for sustaining both people and nature, and the consumption of this limited resource has increased by a factor of 6 since the beginning of the 20th century, which has increased water scarcity in many parts of the world (Obasi, 1997). Now, climate variability and change is posing another threat to the sustainability of this resource. For example, according to Huq et al. (2003), climate change in sub-Saharan Africa may lead to decreased precipitation in semi-arid to arid parts of Africa. Adaptation is dened differently among scholars. According to the Intergovernmental Panel
B *Corresponding author. E-mail: jkashaigili@suanet.ac.tz
CLIMATE AND DEVELOPMENT 1 (2009) 220228

on Climate Change (IPCC) (Bates et al., 2008), climate change adaptation is dened as: initiatives and measures to reduce the vulnerability of natural and human systems against actual or expected climate change effects. Various types of adaptation exist, e.g. anticipatory and reactive, private and public, and autonomous and planned. Bates et al. (2008) dene autonomous adaptations as: those that do not constitute a conscious response to climate stimuli, but result from changes to meet altered demands, objectives

doi:10.3763/cdev.2009.0025 # 2009 Earthscan ISSN: 1756-5529 (print), 1756-5537 (online) www.earthscanjournals.com

Experiences from the Great Ruaha River catchment in Tanzania 221

and expectations which, whilst not deliberately designed to cope with climate change, may lessen the consequences of that change. We contend that it is important to learn lessons from such autonomous adaptations that are widespread in the water sector in order to increase the effectiveness of planned adaptation measures. As to water and adaptation, the IPCC states (Kundzewicz et al., 2007) adaptation to changing conditions in water availability and demand has always been at the core of water management. Consequently this paper seeks to derive lessons from autonomous adaptation that has increased the resilience of a key river catchment and local communities in Tanzania. Resilience (Bates et al., 2008) is dened as: the ability of a social or ecological system to absorb disturbances while retaining the same basic structure and ways of functioning, the capacity for self-organization, and the capacity to adapt to stress and change. According to Adger et al. (2005), adaptation can involve both building adaptive capacity (thereby increasing the ability of individuals, groups or organizations to adapt to changes), and implementing adaptation decisions (i.e. transforming that capacity into action). There is increasing international debate on how best to effectively manage freshwater and adaptation. First, the debate focuses on whether poverty reduction or vulnerability reduction makes for better adaptation (Schipper, 2007). Second, in relation to adaptation across scales (e.g. Adger et al., 2005), discussion focuses on identifying scales of action to deliver more effective adaptation, and how to get actors at different geopolitical scales to work together rather than hinder adaptation work. Third, there is the question of how adaptation measures can best be integrated into society. For example, Ross and Dovers (2008) believe that: the most prominent success factors, barriers and gaps that affect environmental policy integration like climate adaptation relate to leadership, long term embedding of environmental

policy integration capacity.

and

implementation

Considering that climate change impacts are already being felt, interventions aimed at reducing negative impacts and risks are required. Irrespective of the motivation for adaptation, either purposeful or unintentional adaptations can generate short-term or long-term benets (i.e. Adger et al., 2005). This paper presents ndings from an assessment of the adaptation lessons from a WWF project in the Great Ruaha River catchment (GRRC) in Tanzania that was prepared as one of six case studies for a larger review (Pittock, 2009). The research looked at autonomous adaptation in the freshwater sector to derive lessons on what motivated the societies to change, which factors led to more successful adaptation, and how the interventions may best be sustained. The ndings are intended to contribute to the global debate on how better to adapt to climate change.

2. Great Ruaha River catchment and the WWF-Ruaha Water Programme


The Great Ruaha River (GRRC) is a large subcatchment of the Ruji River basin in Tanzania. The Ruji is the largest basin out of the nine hydrological basins in Tanzania, with the drainage area of about 177,000 km2. The GRRC (Figure 1) covers an area of about 83,970 km2 and is home to about 6 million people. It contains the Usangu Plains, which lie at an average elevation of 1,100 m above mean sea level (amsl), located between longitudes 338000 E and 358000 E, and latitudes 88000 S and 98300 S. The plains are surrounded by the Poroto, Kipengere and Chunya mountains (Figure 1), with elevations up to 3000 m amsl. The Usangu wetlands (Western Utengule and Eastern Ihefu wetlands), Selous Game Reserve and Ruaha National Park depend on the waters of the Great Ruaha River (GRR). These ecosystems are of both national and international importance as they are sources of foreign exchange generated through tourism and sport hunting while some

CLIMATE AND DEVELOPMENT

222 Kashaigili, Rajabu and Masolwa

FIGURE 1 Map showing drainage patterns, Usangu wetlands, Ruaha National Park and the Usangu Plains within the GRRC in the Ruji Basin

of the swamps are designated under the Ramsar Convention on Wetlands. Furthermore, the countrys major hydropower plants of Mtera and Kidatu use the waters of the GRR and account for about 48.5% (280 MW out of 577 MW) of installed generation capacity connected to the national power grid. In the GRR headwaters, 46% of the 1.5 million residents live in poverty. The average income is around USD0.80 per day and it is an economy largely based on agriculture. Since the early 1990s the GRR has experienced decreased ows. The GRR is normally perennial, and the successive cessations of dry season ows since 1993 is unprecedented. Since 1957 rainfall

in the lowland portion of the catchment has declined, a trend many fear will be exacerbated by climate change. Furthermore, there has been a change in land use with larger areas converted to agriculture (SMUWC, 2001; Kashaigili et al., 2006). Both the human population and the area under irrigation have expanded, increasing the demand for water as well as conicts among competing users. Many people in the catchment are subsistence farmers depending on rainfed agriculture. The construction of intakes and diversions to abstract water from rivers for supplementary irrigation in order to minimize the risk of crop failure has resulted in the GRR and its tributaries being severely drained.

CLIMATE AND DEVELOPMENT

Experiences from the Great Ruaha River catchment in Tanzania 223

There are many concerns about the GRR drying up (Hirji and Davis, 2009). Such concerns instigated Prime Minister Sumayes announcement in 2001 that the government of Tanzania is committing its support for a programme to ensure that the GRR has a year round ow by 2010. The local concerns mainly arise from the fact that the human population and their livestock depend on land, water and other natural resources available in the catchment to sustain their livelihoods. Their long-term survival depends largely on the sustainable management of the resources of the catchment and on the maintenance of minimum ows in the rivers during the dry season (Kashaigili et al., 2005). The Ruji Basin Water Ofce (a government agency responsible for water resources planning, conservation of water sources and water-based ecosystems, protection of water resources, granting of water rights and conict resolution) is handling the new challenges of regulating demand and supply to efciently allocate a valuable and scarce water resource among competing users. Concerns at the national level arise from the fact that the bulk of the water required for hydroelectric power (HEP) generation at Mtera and Kidatu hydropower plants has its source in the GRRC. Dwindling water supplies from the GRRC, especially during the dry season, negatively affect the existence of important ecosystems such as the Western (Utengule) wetland, Eastern (Ihefu) wetland and Ruaha National Park. Another national concern is food security. Irrigated paddy in Usangu is estimated to produce about 105,000 tons of paddy (equivalent to 66,000 tons of rice) per annum, about 14% of the total annual rice production in Tanzania (Kadigi et al., 2004). In monetary terms annual income from the rice crop is USD15.9 million, which is currently supporting about 30,000 agrarian families in Usangu (Kadigi et al., 2004), with an average income per family of USD530.95 per annum. Despite all that, the challenge remains of how to ensure equitable allocation of available water resources and improve water productivity, while strengthening institutions at the grassroots level and providing alternative livelihoods,

as well as coping with the pressures of a changing climate. By considering the above challenges and concerns, the WWF Tanzania Programme Ofce1 commenced a project in 2003 to promote integrated and sustainable water use and management of natural resources in order to maintain ecosystem functioning for improved livelihoods. The overall objective of the project is that by 2010 the people of the GRRC have the capacity to plan, manage and utilize their water and related natural resources sustainably and, in so doing, alleviate poverty and improve livelihoods. Achievement of sustainable water resource management in the GRRC is to be achieved through integrated capacity building and action at national, catchment, district and community levels by 2010. It was envisaged that the overall and specic objectives of the project would be achieved by ensuring that: An Integrated River Basin Management (IRBM) plan was completed and operational in the GRRC. B Local governments and communities effectively participated in water resource management in line with the national 2002 Water Policy. B Water resources management issues relating to the decreased ows of the GRR were addressed and alternative economic activities that contribute to improved livelihoods were implemented. B Local governments and communities were aware of and understood water resource management and related environmental issues in the GRRC.
B

A key indicator of progress of this project was identied as the restoration of perennial natural ows in the GRR.

3. Methods
This research was commissioned by WWF and undertaken in late 2008 and early 2009, led by Kashaigili (2008). The study did not assess progress

CLIMATE AND DEVELOPMENT

224 Kashaigili, Rajabu and Masolwa

against the projects planned objectives and indicators (above) and instead sought to derive lessons concerning (unplanned) autonomous adaptation to climate change. We applied a largely qualitative analytical framework developed by Pittock (2009) to assess (1) (unplanned) climate change adaptation, (2) socio-economic outcomes and (3) conservation outcomes. The study applied participatory methods (i.e. focus group discussions, direct observation, unstructured interviews) to elicit relevant information. Focus group discussions were conducted with four Community Conservation Banks (COCOBAs) and ve Water User Associations (WUAs) while interviews were conducted with the Ruji River Basin Ofce ofcials, WUA leaders, District ofcials and WWF project implementation staff. Retrospectively, the study assessed the condition before and after WWF intervention to gain an understanding of the effectiveness of the freshwater management and other autonomous adaptation measures and to derive lessons on factors that helped or hindered their implementation. The key measures considered in this project included: the number of established WUAs and COCOBAs; existence of conicts; restoration of headwaters and riparian zones; development of alternative, more environmental friendly income-generating activities; training undertaken on environmental education and entrepreneurship skills; agreements on water scheduling; and the development of water supply solutions.

and river banks, afforestation, reduction of valleybottom (vinyungu) farming and demolition of houses built near water sources. These interventions were achieved through the establishment of local WUAs2 and Apex bodies (an organ overseeing WUAs functions in a watershed) which set and enforced regulations (bylaws). Through WWF, eight WUAs out of ten (Table 1) have been established in eight of the 16 Districts of the GRRC, and more are in the process of establishment. Moreover, construction of Kangaga Dam with a capacity of 40,860 m3 at a cost of USD42,373 has ensured availability of adequate water for livestock as well as for domestic consumption during the dry season and dry years. Consequently, peoples vulnerability to the impacts of drought and low ows in the dry season has been minimized. However, it is not easy to quantify the outcomes for the simple reason that the WWF interventions were not directly aimed at climate change adaptation. As such, insufcient indicators were put in place and monitored in order to quantify the outcomes. The outcomes are largely based on having wellorganized and trained WUAs, COCOBAs,3

TABLE 1 Water User Associations in the GRRC formed by WWF Name of Water User Association Mkoji 20012003 2005/2006 2008 2008 2005/2006 2001/2002 2004 2004 2008/2009 2004/2005 RBMa WWF WWF WWF WWF SMUWC/ RBWOc Halali Nyando Ndembera Mtitu WWF WWF WWF WWF Date established Funding

4. Results 4.1. Adaptation outcomes


The WWF programme has achieved a number of climate change adaptation outcomes, although they were not planned during the design of the programme. For example, over-dependency on water to sustain livelihoods has been greatly reduced as a result of communities being engaged in fewer water-intensive economic activities. Likewise, degradation of the catchment has been reduced by managing the watersheds properly through reduction of grazing near water sources

Mswiswi Mambi Mpolo Chimala MAMREMAb

a Established during the River Basin Management Project and also supported by RIPARWIN. b Mapogoro and Mfumbi Resource Management Association. c Established during SMUWC project.

CLIMATE AND DEVELOPMENT

Experiences from the Great Ruaha River catchment in Tanzania 225

committed communities and supportive village, ward and district governments.

4.2. Livelihood outcomes


Livelihood strategies have diversied from agriculture, brewing and charcoal production into activities requiring less water, notably retailing, manufacturing clothing and bee-keeping. Secure water supplies have supported livestock production, and sh farming in water storages has proved particularly protable. Training 48 rice farmers in better production practices has seen some double their yields. Five Farmer Field Schools specializing in better rice production techniques have been established. Farmers are now better managing the application of water to their paddy elds, from the business-as-usual condition of around 30 cm depth to around 15 20 cm now, doubling water efciency. Moreover, farmers now have an agreed growing calendar indicating a start to the growing season, facilitating improved irrigation scheduling to reduce transmission losses and avoid diverting low ows from the rivers. People who joined the 20 COCOBAs established during the project are now comparatively nancially better off. One COCOBA member said: Most of our fellow villagers who are not members of COCOBAs are becoming poorer and leading a difcult life as compared to COCOBA members. However, there is room for more people to join COCOBAs as the groups have opened doors for more members. Initially, the group membership was limited to 30 persons only. However, the groups can now take as many members as is economically viable and it does not affect the efciency and effectiveness of the group (F. Mwaitegelasye, 2008).

Ihefu wetlands, and the number of zero ows downstream into the National Park has been reduced from almost 3 months per year to less than a month. There is increased awareness of sustainable management and use of water and other catchment resources largely resulting from interventions by WWF, including the Sustainable Management of Usangu Wetlands and its Catchment (SMUWC), a DFID-supported study in 1998 2001, and the Raising Irrigation Productivity and Releasing Water for Intersectoral Needs (RIPARWIN), a river-basin management research project under the support of DFID-KAR and IWMI. Presently, people who have beneted from project interventions are no longer involved in charcoal making because of their understanding of the bylaws that prohibit indiscriminate felling of trees to make charcoal, and are undertaking other activities that do not degrade the environment. A villager from a community that has implemented these measures says of adjacent villages that have not: They are wasting water and they dont know how to conserve water and environment. They grow crops near the water sources, they still make charcoal, start bush res to catch animals, and there are always conicts on water use (Z. Mwakyokola, 2008).

5. Discussion 5.1. Motivations for change


Local communities were keen to implement these adaptations due to their vulnerability to water scarcity and pollution and their need to improve their livelihoods to reduce poverty. Thus, vulnerability has been a catalyst for people to adapt, and this agrees with the argument by Schipper (2007) that we should favour poverty reduction or vulnerability reduction rst for better adaptation. On the other hand, direct interventions to reduce physical vulnerability (e.g. new water storage) have also aided adaptation (Adger et al., 2005). Therefore it can

4.3. Environmental outcomes


The conservation of riparian zones and restoration of springs and river ows are of benet to biodiversity. Flows have recommenced into the

CLIMATE AND DEVELOPMENT

226 Kashaigili, Rajabu and Masolwa

be argued that the two are mutually reinforcing. For the Tanzanian Government, the project has attracted resources to implement its water policy (URT, 2002) in the basin.

5.2. Sustainability and funding


The sustainability of these adaptations, such as enforcement of water rules, depends on ownership and implementation by the local community, which is likely to continue given the strengthened local institutions and livelihood benets derived thus far. This is critical in a developing country like Tanzania where strong and more accountable local institutions are essential to embed sustainable development programmes, given the limited reach of national institutions. This case supports the conclusion of Tompkins and Adger (2004) that community-based management enhances adaptive capacity by building networks that are important for coping with extreme events and by retaining the resilience of the underpinning ecosystems. Funding has so far come from WWF and the EU. There is USD951 million in national and donor funds held by the Tanzanian Government for use up to 2011 to support this type of water sector development nationally. The governments intention to foster this kind of river basin management through a new water law and allocation of central funds and fees from water users to local management institutions (URT, 2002) is yet to be realized, jeopardizing the long-term sustainability of these adaptations. Furthermore, the Tanzanian Government has a policy of expanding irrigation (URT, 1997), which if implemented poorly in the GRRC, may further reduce river ows. A further sustainability question is whether the adaptations implemented are sufcient to restore perennial ows to the GRR to achieve one of the programmes main objectives, and whether these measures may prove insufcient with further climate change. WWF and colleagues have recently completed a review of environmental ow assessment for the GRR to determine

the water required to meet particular environmental and social objectives, and options for achieving them. This concluded that the better catchment management measures applied, while highly benecial in increasing river ows and reducing vulnerability of local communities to water scarcity in the upper basin, would not exceed the threshold needed to pass dry season water ow through the Usangu wetland and further downstream. Consequently the diversion of an upstream tributary, the Ndembera River, around the wetland has been proposed to provide a base ow to the GRRs main stream (Mwaruvanda et al., 2009). Although the additional water supply that can be generated by the catchment restoration appears limited, it has provided benets and bought time to consider further adaptation measures, such as the proposed diversion.

5.3. Barriers and lessons


Work with government agencies locally was hampered as newly trained ofcers took up better employment offers elsewhere. Lessons for successful adaptation from the programme are: that seed funding is essential for the transition; improvements in livelihoods motivates change; establishing and strengthening local institutions and making links to basin and national institutions make this change more sustainable, while meeting the criteria of Adger et al. (2005) (effective, efcient, equitable and legitimate). However, while good progress has been made locally using foreign funds and the National Water Policy currently provides legitimacy, the lack of implementation of national nancing measures for the WUAs is neither effective nor efcient. Further, there is no evidence of the processes of the UN Framework Convention on Climate Change inuencing governmental actions in the GRRC. As outlined above, the measures being employed by WWF to date are not exhaustive and new actions will have to be considered. For example, addressing water scarcity in the

CLIMATE AND DEVELOPMENT

Experiences from the Great Ruaha River catchment in Tanzania 227

catchment requires the understanding of the whole water balance and not just blue water, which has been the focus of management interventions to date. Furthermore, a study is suggested on how to improve water productivity, which is one of the key strategies for improving rain-fed irrigation and at the same time reducing the downstream impacts.

freshwater and adaptation. Thus, as for Schipper (2007), climate change adaptation will be facilitated by a focus on sustainable development and vulnerability reduction, with an explicit integrated approach that accounts for factors such as poverty reduction and rural livelihoods.

Acknowledgements
The authors are very grateful to nancial support from WWF-TPO which made this study possible. Special thanks to the communities under the Ruaha Water Programme for their cooperation. We are also very grateful to Jamie Pittock, John Matthews and anonymous reviewers for comments. This version remains exclusively our responsibility. Dr Kossa Rajabu died after this paper was drafted. We acknowledge his considerable contribution and extend our sympathies to his family. We can ill afford to lose such talented experts.

5.4. Potential to scale up


The adaptation measures implemented so far were technologically simple and decentralized, and applied existing knowledge, supporting Dovers (2009) proposition that societies can make great progress towards climate adaptation by implementing what is already known. This approach to adaptive catchment management could be scaled up, given its modest cost and the national and donor funds available in Tanzania and other developing countries. For instance, the measures have been applied in only half of the 16 districts in the GRRC thus far.

Notes
1. The project received nancial support from WWFUK and the European Union (EU), and worked in collaboration with the Tanzanian Ministry of Water and Irrigation (through the Ruji Basin Water Ofce) and the District Councils of Mbarali, Mbeya Rural, Chunya, Mundi, Makete, Njombe, Kilolo and Iringa Rural District. 2. The cost of forming one WUA is around TSH30,000,000 (USD24,000). One WUA on average is composed of 20 villages. WUAs were trained on water policy and laws, water management, integrated water resources management concepts, conicts management, group management, data management, banking, leadership, environmental education and preparation of management plans. 3. The costs of forming one COCOBA is around TSH12,500,000 (USD10,000).

6. Conclusions
The inexpensive, grass-roots climate adaptation measures applied in the Great Ruaha demonstrate how incremental action to restore ecosystem functions and better manage natural resources can increase resilience to water scarcity. It highlights the importance of strengthening the capacities of local people and organizations to improve governance, diversify the local economy and institute adaptive management practices. This case also emphasizes the need for governments to support local organizations with appropriate mandates and nancial independence to undertake ongoing adaptive management. It can therefore be concluded that strengthening local institutions and building capacity among others are fundamental to adaptation and sustainable freshwater management and should be the focus for future policy and interventions. The study also contributes to an international debate on how to manage

References
Adger, W. N, Arnell, N. W. and Tompkins, E. L., 2005. Successful adaptation to climate change across

CLIMATE AND DEVELOPMENT

228 Kashaigili, Rajabu and Masolwa

scales. Global Environmental Change Part A, 15(2). 77 86. Bates, B. C., Kundzewicz, Z.W., Wu, S. and Palutikof, J. P. (eds), 2008. Climate Change and Water. Technical Paper of the Intergovernmental Panel on Climate Change. IPCC Secretariat, Geneva. Dovers, S., 2009. Normalizing adaptation. Global Environmental Change, 19(1). 4 6. Hirji, R. and Davis, R. (eds), 2009. Environmental Flows in Water Resources Policies, Plans, and Projects. Case Studies. The World Bank, Washington, DC. 41 49. Huq, S., Rahman, A., Konate, M., Sokona, Y. and Reid, H., 2003. Mainstreaming Adaptation to Climate Change in Least Developed Countries (LDCs). www. un.org/special-rep/ohrlls/ldc/LDCsreport.pdf. Kadigi, R. M. J., Kashaigili, J. J. and Mdoe, N. S., 2004. The economics of irrigated paddy in Usangu basin in Tanzania: water utilization, productivity, income and livelihood implications. Physics and Chemistry of the Earth, 29. 10911100. Kashaigili, J. J., 2008. Impacts of land-use and landcover changes on ow regimes of the Usangu wetland and the Great Ruaha River, Tanzania. Physics and Chemistry of the Earth, 33. 640647. Kashaigili, J. J., Kadigi, R. M. J., Lankford, B. A., Mahoo, H. F. and Mashauri, D. A., 2005. Environmental ows allocation in river basins: exploring allocation challenges and options in the Great Ruaha basin, Tanzania. Physics and Chemistry of the Earth, 30(1116). 689 697. Kashaigili, J. J., Mbilinyi, B. P., McCartney, M. and Mwanuzi, F. L., 2006. Dynamics of Usangu Plains wetlands: use of Remote Sensing and GIS as management decision tools. Physics and Chemistry of the Earth, 31. 967 975. Kundzewicz, Z. W., Mata, L. J., Arnell, N. W., Doll, P., Kabat, P., Jimenez, B., Miller, K. A., Oki, T., Sen, Z. and Shiklomanov, I. A., 2007. Freshwater resources and their management. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, M. L.

ParryJ. P. Palutikof, P. J. van der Linden and C. E. Hanson (eds). Cambridge University Press, Cambridge, UK. 173 210. Mwaitegelasye, F., 2008. Personal communication. Member, Mkombozi Community Conservation Bank, 20 March. Mwakyokola, Z., 2008. Personal communication. Secretary, Mbuyuni Intake Water Users Association, 19 March. Mwaruvanda, W., Kashaigili, J. J., OKeeffe, J. H. and Von Der Heyden, C., 2009. Restoring ows in the Great Ruaha River, Tanzania. International Conference on Implementing Environmental Water Allocations. Port Elisabeth, South Africa. Obasi, G. P., 1997. Climate Change and Freshwater Management. World Meteorological Organization. First World Water Forum. Marrakech, Morocco. Pittock, J., 2009. Lessons for climate change adaptation from better management of rivers. Climate and Development, 1(3). 194 211. doi:10.3763/cdev.2009.0021. Ross, A. and Dovers, S., 2008. Making the harder yards: environmental policy integration in Australia. Australian Journal of Public Administration, 67(3). 245260. Schipper, E. L. F., 2007. Climate Change Adaptation and Development: Exploring the Linkages. Tyndall Centre Working Paper No. 107. Tyndall Centre for Climate Change Research, Norwich, UK. SMUWC (Sustainable Management of the Usangu Wetlands and its Catchment), 2001. Final Report, Water Resources. Supporting Report 7, Vol. 3, Directorate of Water Resources, Dar es Salaam, Tanzania. Tompkins, E. L. and Adger, W. N., 2004. Does adaptive management of natural resources enhance resilience to climate change? Ecology and Society, 9(2). 10. URT (United Republic of Tanzania), 1997. The Agricultural and Livestock Policy. 161. www.tzonline.org/ pdf/agricultureandlivestockpolicy.pdf. URT (United Republic of Tanzania), 2002. The National Water Policy. Ministry of Water and Livestock, Dar es Salaam, Tanzania.

CLIMATE AND DEVELOPMENT

case study

Adapting to climate change in the Godavari River basin of India by restoring traditional water storage systems
BIKSHAM GUJJA1, SRABAN DALAI2, HAJARA SHAIK3 and VINOD GOUD4,*
1 2

Senior Policy Advisor, Global Freshwater Programme, WWF-International, CH-1196, Gland, Switzerland GIS Coordinator, ICRISAT-WWF International Project, Patancheru, Andhra Pradesh, India 3 Sociologist and independent consultant, Gland, Switzerland
4

Project Coordinator, ICRISAT-WWF International Project, Patancheru, Andhra Pradesh, India

Food, climate and water crises are interrelated and interdependent. Climate change is projected to signicantly alter rainfall patterns, with implications for the annual runoff for the Godavari River basin of the Indian subcontinent. Agriculture, especially rain-fed agriculture, will be particularly affected, due to changes such as periodicity and intensity of rainfall. This paper describes eld interventions in 20052007 designed to restore traditional water management systems (in the form of water tanks; that is, wetlands embedded in a semi-arid region), with the aim of mitigating the effects of increased climate variability and the frequency of weather extremes. Our ndings suggest that traditional water management methods can be both socially and economically effective in coping with variability in precipitation patterns, decentralizing management institutions, improving crop productivity and increasing groundwater recharge. This approach is preferable to large projects for increasing water storage capacity or expanding areas under irrigation, which are expensive, and can displace people and degrade ecosystems. Keywords: climate change; groundwater; restoration; tanks; water productivity

1. Introduction
Indias water needs are expected to grow dramatically by 2050, with demand increasing beyond current supplies. According to current estimates, even if agricultural demand per capita does not increase, domestic demand will increase by 45% between 2000 and 2050, and total demand will increase by 65% (NCIWRD, 1999; Planning Commission, 2007). Therefore India needs to secure additional water supplies if it is to avoid a major water crisis. Indias national water management policies promote integrated water management at the river-basin level (Palanisami, 2006). However, implementing such policies is often challenging or impractical. In India, state governments control the development and management of
B *Corresponding author. E-mail: v.goud@cgiar.org
CLIMATE AND DEVELOPMENT 1 (2009) 229240

water infrastructure, but many river basins are shared by more than one state. A further challenge is that water allocations often exceed the total mean ow, leading to conicts between stakeholders (Gujja et al., 2006a,b,c). Climate change is likely to add further complexity to the challenge of water management in India. The impact of climate change at the basin level is difcult to predict, but there are some general predictions about the impact on water resources (Gosain et al., 2006). Increasing temperatures could cause signicant shifts in the variability of river runoff, and a 18C increase in mean annual temperature could result in a 15% reduction in inows (Cai and Whetton, 2001; Cai and Cowan, 2008). Efforts to manage water rarely focus on improving the livelihoods of poor people or conserving

doi:10.3763/cdev.2009.0020 # 2009 Earthscan ISSN: 1756-5529 (print), 1756-5537 (online) www.earthscanjournals.com

230 Gujja et al.

freshwater ecosystems. India is investing heavily in water infrastructure to improve supplies, but often grossly overestimates water availability and fails to allow for increased variability due to climate change. According to several estimates, India will need another 120 major storage dams by 2050 to meet projected water needs (NCIWRD, 1999; CWC, 2005 2006). Building additional dams will create conict, displace people, degrade ecosystems and increase the cost of water delivery. Although methods such as improving efciency in water use will play an important role in resolving Indias water problems, it is clear that it will also be necessary to increase water availability. The challenge is to devise practical methods that improve water availability and meet current demands, while providing enough water to meet future demand in the context of uncertainty due to climate change. This paper looks at the possibility of improving water availability in the Maner sub-basin by restoring traditional water tanks (i.e. wetlands embedded in a semi-arid region), instead of developing new infrastructure. The sub-basin is a more practical focus for study than a basin, because all of the subbasins catchment areas are in the same state: Andhra Pradesh.

2. Traditional water tanks in the Maner sub-basin


People in Andhra Pradesh state have been building water tanks, known as cheruvu (big tank) or kunta (small tank), for over 2,000 years, and there are about 74,000 water tanks across the state, according to the World Bank (2007). Some of the water tanks are still functioning more than 1,200 years after they were built. Water tanks are multifunctional, and their uses vary with climate and rainfall. Until recently, tanks were the main source of drinking water for humans and livestock. They are also a source of sh (and thus host a wide range of birds), have a ceremonial role for local communities, are the main source of groundwater recharge and have

traditionally been used to manage scarce and erratic rainfall. The Maner sub-basin is located in the semi-arid mid-Godavari River basin (Figure 1), where most rain falls during the monsoon (JuneOctober). Mean annual rainfall normally varies from 629 to 1,391 mm, although long-term data show years of very low rainfall (e.g. 517 mm in 1920 and 598 mm in 1972) and of very high rainfall (e.g. 1,391 mm in 1983) (Mitchell and Jones, 2005). For agriculture, the total annual rainfall is less important than its timing and intensity, both of which are likely to be seriously affected by climate change (Bates et al., 2008). In the Maner sub-basin, 38% of the population work in agriculture as farmers and a further 40% are employed in agricultural services (such as eld labour). About 35% of the sub-basin is used for agriculture; the net cultivated area is 455,000 ha, and 127,000 ha is forest (Figure 2). The sub-basin also includes 1,761 towns and large villages, including two district headquarters: Warangal and Karimnagar (Directorate of Economics and Statistics, 2003, 2005, 2006 2007). Irrigation is central to agriculture in the Maner sub-basin, traditionally through the use of water tanks. About 44% of the total cultivated area is irrigated, of which 22% (74,000 ha) is irrigated using surface water; 39,000 ha of this area is irrigated using water tanks, and the remainder is irrigated by canals and groundwater (which is recharged through water tanks). The Maner sub-basin is divided into 24 micro-basins, all of which have water tanks, and most of which use tank irrigation. It is difcult to estimate the exact amount of water used in the Maner sub-basin, but the agricultural water use has been estimated based on production of major irrigated crops (Table 1). The total water used for the four major crops is 1,621 million m3. The major crops (rice, maize, chillies and cotton) account for more than 85% of the value, production and cultivated area. Rice is the main irrigated crop; it occupies more than 46% of the total irrigated area and accounts for more than 75% of total water used in the Maner sub-basin. The total water use in irrigated agriculture is around 2,000 million m3, which

CLIMATE AND DEVELOPMENT

Restoring traditional water storage systems in the Godavari River basin, India 231

FIGURE 1 Maner sub-basin Source: WWF.

represents about 17% of the total rainwater in the basin. In the Maner sub-basin, it will be a major challenge to adapt to the effects of climate change and improve land productivity using local water resources without degrading the ecosystem.

3. Restoration of water storage tanks


Despite their importance, water tanks are declining in number, size and capacity (Vaidyanathan, 2001). This decline is partly due to changes in the way that water tanks are managed, which

FIGURE 2 Percentage of land use in Maner sub-basin districts and Andhra Pradesh

CLIMATE AND DEVELOPMENT

232 Gujja et al.

TABLE 1 Water consumption by major crops Crop Litres of water used per kg produced Rice Maize Cotton Chillies
Source: Hoekstra and Chapagain, 2004.

2,656 450 5,300 5,300

reects a change from traditional governance structures to governance by state government departments. In the past, particularly before the Land Reforms Act (Chandra Pal, 1989), poor people and low-caste communities had little or no land to irrigate, but they were responsible for managing the water and cleaning the water tanks and canals. Clearly it is not desirable or necessary to recreate these past social conditions for water tank maintenance. However, it is

important to acknowledge that water tank restoration and management is complex, and must adapt traditional understanding of the ecosystem to the socio-political situation of present-day India. Previous programmes to restore water tanks have been unsystematic, and have considered neither the ecosystem nor local hydrology. Today, the Maner sub-basin has only 6,234 water tanks, which cover about 5% of the catchment area (Figure 3). On average, there is one tank for every 600 people. Most water tanks are small (1 10 ha), although some are large (.100 ha), and there are also the new lower and upper Maner reservoirs which span over 6,040 ha. Each hectare of water tank supports about 5.6 ha of irrigated area (including areas irrigated using groundwater). The primary data sources used in the research reported here to delineate the watershed, subwatersheds and tanks, and to assess the water

FIGURE 3 Maner catchment showing micro-basins and major water tanks Source: WWF.

CLIMATE AND DEVELOPMENT

Restoring traditional water storage systems in the Godavari River basin, India 233

storage potential of the tanks, were Survey of India topographic maps at a scale of 1:50,000, multispectral band Landsat thematic mapper satellite imagery, and digital elevation model. The geographic information system (GIS) and mapping software used in this study were ArcGIS 9.x, and handheld global positioning system for eld sample collection and verication. The current storage capacity of many of the water tanks in the Maner sub-basin is far less than their potential storage capacity, because many are lled with silt. De-silting the tanks would greatly increase the overall storage capacity, providing a more secure water resource to meet current and future water demands. If all tanks were de-silted and renovated, would there be sufcient water through rainfall, particularly during drought years, to ll the renovated tanks? Average annual rainfall over the basin in 2001 2005 was about 973 mm. This amounts to about 12 billion m3 of rainwater falling over the basin. If all the water tanks in the Maner sub-basin were de-silted to 3 m, they could store 1.764 billion m3 of rainwater. It could be increased further to 2.940 billion m3 if the tanks were dug to an average of 5 m. Thus, the water storage of the tanks could be one-fourth to onesixth of the average rainfall over the basin. In years when rainfall is too low to ll all the tanks, groundwater recharge from de-silted tanks could meet the requirements during that year. Thus, the Maner sub-basin would have enough water to ll the de-silted water tanks without greatly altering river ow. WWF supported a pilot project to renovate traditional water tanks in the Maner sub-basin. The project was implemented by the Warangal-based nongovernmental organization Modern Architects for Rural India (MARI) from March 2005 to February 2007 (WWF-ICRISAT, 2007). The Sali Vagu micro-basin was the area selected for water tank restoration (Figure 4). About 20% of the total irrigated area in the Sali Vagu micro-basin is irrigated from water tanks and the remainder is irrigated with groundwater (which depends on the water tanks for recharge). Twelve traditional water tanks were selected for de-silting

through community participation methods. The main objectives of this project were to:
B

initiate the eld projects to increase the water storage capacity of selected water tanks by de-silting; demonstrate that improved water management can increase the productivity of irrigated crops and reduce the dependency on chemical fertilizers; examine whether restoring traditional water tanks is nancially viable compared to major water infrastructure projects on the same river basin; identify policy changes needed to implement approaches at the watershed level rather than undertaking major water infrastructure projects; develop policy tools for promoting large-scale water tank restoration as an effective, ecologically sound method of improving water management without compromising food security and economic prosperity.

De-silting is the most important aspect of watertank restoration: it increases the capacity of the water tank, provides an organic fertilizer (in the form of silt), and can provide a source of income for local people employed to apply the silt to the elds. The International Crops Research Institute for the Semi-Arid Tropics (ICRISAT) analysed the sediment and found that it contained 6070% clay, and was rich in organic carbon and minerals. The project removed 73,000 tonnes of sediment from the 12 water tanks covering an area of 11 ha and serving 42,000 people, with direct costs of INR1.1 million (USD28,000) and indirect costs (e.g. management) of about INR15 (USD0.30) per tonne. Some benets of the project were: crop productivity increased by INR5.8 million per year; the crops with the highest productivity increases were groundnut and maize; B farmers irrigated about 900 ha more land after the de-silting project;
B

CLIMATE AND DEVELOPMENT

234 Gujja et al.

FIGURE 4 Sali Vagu micro-basin Source: WWF.

CLIMATE AND DEVELOPMENT

Restoring traditional water storage systems in the Godavari River basin, India 235

B B

B B

annual productivity of the sub-catchment increased by INR1.0 million, partly due to the project; groundwater use decreased; partly because the water tanks provided sufcient water for irrigation without using groundwater as a supplement; water pump use decreased, saving electricity; rural employment improved due to increased agricultural activity; a survey quantied this improvement as up to INR500,000 (USD12,000) in additional wages in these tanks; net prot from shing increased by INR160,000 (USD3,000) per year; this increase is expected to last for at least 5 years; habitat for birds improved, and other ecological benets were observed; indirect benets that could be partly attributed to de-silting included reduced numbers of crop pests, decreased use of inorganic fertilizers, reduced migration to cities, and improved fodder availability, resulting in increased milk production.

during the seasons when less agriculture work was available. This pilot project has clearly indicated that the de-silting of tanks on a large scale could result in:
B B

B B B B

tangible benets for crop cultivation and sheries; indirect benets, such as increased B fodder production B organic manure production B use of silt to improve soil in agricultural elds B domestic water; new employment, which decreases migration to cities; signicant resolution of water conicts; reduced need for large-scale water infrastructure projects; large-scale restoration of wetlands.

4. Planning for climate change in the Maner sub-basin


Climate change is likely to affect water availability, agricultural productivity, and terrestrial and aquatic ecosystems in the Maner sub-basin. However, specic changes are difcult to predict, so water management strategies must consider worst-case scenarios and design adaptive solutions accordingly (Milly et al., 2008). The Intergovernmental Panel on Climate Change projections (IPCC, 2007) for freshwater resources suggest that climate change will cause: variations in periodicity and intensity of rainfall; B decreases in water resources in some areas and increases in others; B longer periods of dry season ows in rivers.
B

The nancial benets of the project not only recovered the project costs but created some prot. Extensive surveys and observations also demonstrated that the project had both ecological and agricultural benets. The project used incentives to mobilize communities to participate, and the result was a water tank restoration project based on community effort and available resources. Engaging the community ensured that they reaped direct benets. For example:
B

B B

silt was applied to the elds of 884 farmers, covering an area of 602 ha; most of these farmers reported increased land productivity and reduced costs; increased area and depth of the water in the tanks facilitated improved sh production; sixteen mounds (islands) of soil were created in the water tank area to provide safe habitat for birds; establishing such mounds also reduced the cost of de-silting; the de-silting also provided direct employment to the rural community, particularly

Quantitative projections of changes in precipitation, river ows and water levels at the riverbasin scale are uncertain, but overall, even in areas where precipitation is projected to increase, water resources are likely to become more scarce.

CLIMATE AND DEVELOPMENT

236 Gujja et al.

More research and data analysis are required to make concrete projections, but existing information shows that: climate change will alter rainfall, soil moisture and river ows in ways that are detrimental to agriculture, with changing rainfall patterns disturbing cropping patterns; B global or continental projections of climate change-induced variability are possible, but precise projections at the basin, sub-basin, or watershed levels will take longer to generate, or may prove impossible to generate with high condence; however, these smallerscale areas are where prevention and adaptation initiatives must take place; B estimates and suggestions based on the available historical data for various regions, combined with more local and long-term data, can offer some direction when planning and initiating interventions; B there are no high-condence, India-specic projections of how climate change will affect water resources, agricultural productivity, oods and water infrastructure in coming decades.
B

Lack of data and high-condence projections makes planning difcult for adaptation and economic development initiatives, especially with regard to changing precipitation patterns. Indian peninsular rivers do not receive inows from snow melt and depend entirely on rainfall which, as previously discussed, is highly variable. Annual runoff comes from the few days of

intensive rainfall per year. In the Godavari River basin, the river is dry most years between December and July, with 85% of the discharge occurring between July and October. The smaller the river basin, the shorter and more intense the surface ows. For the Maner sub-basin, downscaled climate projections indicate that annual runoff may shift from 7.6% to 59.9% in coming decades, with a cross-model average of 24% (P.C.D. Milly, pers. commun., February 2008). Even the increased ow in the Maner sub-basin may not reduce water shortages or water conicts, because the rain may fall in a very short period. Unless proper adaptive measures are in place, even increased precipitation and the associated increased runoff, if any, may not actually make more water available. To adapt to climate change, many options need to be explored. Simply increasing large-scale storage through big infrastructure projects may not resolve the water conicts. There are already large variations in the Maner sub-basin ow from year to year (CWC, 2003 2004; Figure 5). Interestingly, the maximum annual ow (which occurred in 1983) was 16.5 times the minimum annual ow (which occurred in 1985), while the rainfall for those years varied by a factor of 2. Although there is a direct relation between precipitation and river runoff, often the intensity and time of rainfall determines river ow. For example, in 1983, the average rainfall was 1,391 mm, the total rainfall over the basin was about 18.2 billion m3 and the outow was 6.2 billion m3, suggesting that river runoff was about 34% of total rainfall. This is comparable

FIGURE 5 Variation in annual discharge and rainfall over the Maner sub-basin

CLIMATE AND DEVELOPMENT

Restoring traditional water storage systems in the Godavari River basin, India 237

to the theoretical calculation of 40% for an average catchment (Shanmugham and Kanagavalli, 2005). However, when rainfall is below average, the river ow drastically reduces. In 1985, average rainfall was 680 mm, total rainfall over the basin was about 7.90 billion m3, and the outow was 0.38 billion m3, suggesting that river runoff was just 5% of the total rainfall far less than the 14% expected from theoretical calculations. This clearly indicates that, at sub-basin level, changes in rainfall due to climate change are likely to result in even more drastic uctuations in river runoff. Based on these actual observations and on longterm data analysis, the following observations can be made about the relationship between rainfall and Maner sub-basin discharges: when rainfall is between 900 mm and 1,000 mm, the runoff is around 15% of the rain; in contrast, at 600 mm, the runoff is less than 5% of rainfall; high rainfall (about 1,300 mm) could result in runoff of more than 34%; if rainfall is above 1,000 mm for two consecutive years, the third years runoff will be proportionally higher, possibly due to soil saturation and reduced evaporation; rapid and high rainfall will result in a sudden increase in runoff, but will not lead to sustainable river ows.

water supplies and adapt to climate change will be essential to protect livelihoods.

5. Discussion
The following questions need to be answered to facilitate water management at the sub-basin level:

5.1. Can increased water demand be met by managing water at the sub-basin level?
Increased water demand can be met, but only with some radical changes to the existing infrastructure of traditional water tanks, open wells and tube wells. More than 80% of the Maner subbasins 434,000 ha of agricultural land is irrigated using groundwater, which relies on water tanks for recharge, so improving maintenance will be a critical part of protecting declining reserves. In terms of agricultural productivity, there is little scope for increasing the net area of cultivation, so to increase production it will be necessary to: improve water availability to the irrigated area; B improve water productivity; B change cropping patterns, shifting to highervalue crops.
B

B B

The population within the Maner sub-basin is expected to increase by 45% (to 5.5 million) by 2050, parallelling the projected national increase. If water consumption increases (NCIWRD, 1999; Planning Commission, 2007) as estimated in national level projections (to 735 m3 per capita), the total water requirement in the Maner sub-basin would be around 4,000 million m3 almost double current estimated water use. While this prediction of demand is on the high side, at least 3,500 million m3 of water will be required, including 2,000 million m3 for agriculture. Given the economic importance of agriculture in the Maner sub-basin, initiatives that help farmers secure

Renovating and de-silting water tanks would greatly increase their capacity. If all of the water tanks in the Maner sub-basin were de-silted to 3 m, they could store 1,764 million m3; if they were de-silted to 5 m, they could store 2,940 million m3. This additional storage could help increase the water supply to meet demand.

5.2. Could water tanks help to manage the impacts of climate change?
When planning for climate change, the most pressing issue may be the threat of extremely severe droughts. Water tank restoration provides an opportunity to recharge groundwater during

CLIMATE AND DEVELOPMENT

238 Gujja et al.

high-rainfall years. This water could be used during times of drought, although droughts lasting longer than two years would certainly create severe water shortages. During such droughts, an existing large storage dam above the Godavari River basin could be used to ll up the water tanks, provided the drought does not span the entire basin and evaporation does not sap the reservoir. There is clearly a limit to the amount of water the water tanks could provide in the face of extreme weather events of long duration. However, water tank restoration is an excellent strategy to help manage annual rainfall uctuations, to store more water to reduce the impact of low rainfall, and to reduce the risks of drought and speed recovery.

5.5. What would be the costs of meeting increased water demand locally?
All initiatives for improving water supplies must be affordable and cost-effective. The cost of removing silt from water tanks is around INR15 per cubic metre, giving a cost of INR25.5 billion (USD635 million) for 1,700 million m3. The de-silting process would need to continue for at least 5 years. A strategic approach towards improving the traditional water systems could avoid large-scale water projects (Gujja et al., 2006c) and also provide direct employment to rural people.

6. Conclusions 5.3. What other options are available for improving agricultural productivity?
Irrigation improves productivity, but high productivity could be achieved using much less water, such as using methods like system of rice intensication (SRI) (Gujja et al., 2007). WWF is working with farmers to implement this method. Results indicate that it is possible to increase productivity from 2.7 tonnes per ha to 3.5 5 tonnes per ha without increasing the irrigated area or water supply. To achieve such results, new approaches and investments in training are needed. An added advantage of the SRI method is that it could reduce methane output from the rice elds by 50% compared to conventional cultivation through ooded irrigation. This paper has argued that renovating traditional water tanks would be a cost-effective, environmentally friendly and socially equitable method of responding to the existing water crisis, adapting to climate change and improving wetland ecosystems in rural India. There are 208,000 village water tanks across India (Vaidyanathan, 2001), meaning that this project could be scaled up to a national level. In the Maner sub-basin, which this paper has examined in detail, water tanks support more than 80% of local water needs and have the potential to store more water and efciently recharge groundwater. Groundwater resources are likely to increase in importance, because they can be a tool to combat the predicted increased frequency and severity of droughts due to climate change. Although water tanks are currently used across the Maner sub-basin, they have become degraded and are underused. We suggest that the water demands of the increasing population can be met by managing water at the sub-basin level. By investing around USD100 per capita (USD550 million for the Maner sub-basin), the water crisis could be substantially resolved while simultaneously creating a vibrant rural economy and functional wetlands. If the sediment from de-silting is used as a fertilizer, investments could be recovered through

5.4. Is it possible to meet increased water demand while improving terrestrial and aquatic ecosystem health?
Being wetlands, water tanks provide signicant habitat for wildlife. De-silting will make water tanks a more productive ecosystem, improving sh habitat (and commercial sh production) and attracting migratory birds, without signicantly reducing downstream environmental ows.

CLIMATE AND DEVELOPMENT

Restoring traditional water storage systems in the Godavari River basin, India 239

increased crop production. Intensication methods such as SRI could help to meet future cereal requirement by changing farm practices and using groundwater resources more effectively, which might even contribute to the mitigation of future climate change. Overall, the restoration of water tanks in the Godavari River basin presents a unique opportunity to improve water resources while simultaneously improving the local economy and ecosystems.

Acknowledgements
We thank Dr P. C. D. Milly for providing the data on variability of river runoff with climate change in India. We also thank MARI for providing the data on its water tank restoration project in Sali Vagu, funded by the WWF-ICRISAT project.

References
Bates, B. C., Kundzewicz, Z. W., Wu, S. and Palutikof, J. P. (eds), 2008. Climate Change and Water. Technical Paper of the Intergovernmental Panel on Climate Change. IPCC Secretariat, Geneva. Cai, W. and Cowan, T., 2008. Evidence of impacts from rising temperature on inows to the Murray Darling basin. Geophysical Research Letters, 35. L07701. doi:10.1029/2008GL033390. Cai, W. and Whetton, P. H., 2001. A time-varying greenhouse warming pattern and the tropical-extratropical circulation linkage in the Pacic Ocean. Journal of Climate, 14. 33373355. Chandra, Pal., 1989. Legal aspects of land reforms. Land reforms in India: Achievements, Problems and Prospects, M. L. Sharma and R. K. Punia (eds). Ajanta Publications, Delhi. CWC (Central Water Commission), 2003 2004. Annual Report. CWC, New Delhi. CWC (Central Water Commission), 2005 2006. Annual Report. CWC, New Delhi. Directorate of Economics and Statistics, 2003. Glimpses of the District 2002 03: Karimnagar, Medak, Nizamabad and Warangal Districts. Directorate of Economics and Statistics, Government of Andhra Pradesh, Hyderabad. Directorate of Economics and Statistics, 2005. Statistical Abstract of Andhra Pradesh 2005. Directorate of Economics and Statistics, Government of Andhra Pradesh, Hyderabad.

Directorate of Economics and Statistics, 2006 2007. An Outline of Agricultural Situation in Andhra Pradesh. Directorate of Economics and Statistics, Government of Andhra Pradesh, Hyderabad, India. Gosain, A. K., Rao, S. and Debajit, B., 2006. Climate change impact assessment on hydrology of Indian River basins. Current Science, 90(3). 346 353 Gujja, B., Pangare, G. and Alagh, Y. K., 2006a. Interlinking of Rivers in India. NCSCILR and Academic Foundation, New Delhi. Gujja, B., Joy, K. J., Paranjape, S., Goud, V. and Vispute, S., 2006b. Million revolts in the making, water conicts in India. Economic and Political Weekly, 41(7). 18 24. www.epw.org.in. Gujja, B., Ramakrishna, S., Goud, V. and Sivaramakrishna, 2006c. Perspectives on Polavaram A Major Irrigation Project on Godavari. Academic Foundation, New Delhi. Gujja, B., Riddell, P., Goud, V., Dalai, S., Holland, R. and Rupela, O. P., 2007. More Rice with Less Water: SRISystem of Rice Intensication. New Concept Information Systems, Hyderabad, India. Hoekstra, A. Y. and Chapagain, A. K., 2004. Water Footprints of Nations: Value of Water Research. Report Series No. 16, UNESCO-IHE, Delft, The Netherlands. www.waterfootprint.org. IPCC (Intergovernmental Panel on Climate Change), 2007. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, M. L. Parry, O. F. Canziani, J. P. Palutikof, P. J. van der Linden and C. E. Hanson (eds). Cambridge University Press, Cambridge, UK. Milly, P. C. D., Betancourt, J., Falkenmark, M., Hirsch, R. M., Kundzewicz, Z. W., Lettenmaier, D. P. and Stouffer, R. J., 2008. Stationarity is dead: whither water management? Science, 319. www.sciencemag. org. Mitchell, T. D. and Jones, P. D., 2005. An improved method of constructing a database of monthly climate observations and associated high resolution grids. International Journal of Climatology, 25. 693 712. NCIWRD (National Commission for Integrated Water Resources Development), 1999. Ministry of Water Resources of India: Integrated Water Resource Development A plan for Action. NCIWRD, Government of India, New Delhi. Palanisami, K., 2006. Sustainable management of tank irrigation systems in India. Journal of Developments in Sustainable Agriculture, 1. 34 40. Planning Commission, 2007. Report of the Working Group on Natural Resources Management, Eleventh

CLIMATE AND DEVELOPMENT

240 Gujja et al.

Five Year Plan (20072012). Vol. I: Synthesis. Planning Commission, Government of India, New Delhi. Shanmugham, C. R. and Kanagavalli, J., 2005. Technology of Tanks (Practitioners Handbook). DHAN Foundation, Madurai, India. Vaidyanathan, A. (ed.), 2001. Tanks of South India. Centre for Science and Environment, New Delhi. World Bank, 2007. Andhra Pradesh Community-Based Tank Management Project, Report No.: AB2774.

Command Area Development Authority, Irrigation and Command Area Development Department, Government of Andhra Pradesh, India. WWF-ICRISAT, 2007. Community Action Based Policy Dialogue on Restoration and Sustainable Management of Tanks in the Mid Godavari Basin. Consolidated WWF-ICRISAT Project Completion Report, Modern Architects for Rural India (MARI), Warangal, Andhra Pradesh, India.

CLIMATE AND DEVELOPMENT

case study

Freshwater management and climate change adaptation: Experiences from the central Yangtze in China
XIUBO YU1, LUGUANG JIANG1, *, LIFENG LI2, JINXIN WANG3, LIMIN WANG4, GANG LEI4 and JAMIE PITTOCK5
1

Institute of Geographic Sciences and Natural Resources Research, Chinese Academy of Sciences, 11A Datun Road, Chaoyang District, Beijing 100101, China

2 3

WWF International, Avenue du Mont-Blanc 27, 1196 Gland, Switzerland School of Urban and Environmental Science, Xuzhou Normal University, 101 Shanghai Road, Xuzhou 221116, China 4 WWF China Programme Ofce, Room 1609 Wen Hua Gong, Beijing Working Peoples Culture Palace, Beijing 100006, China 5 Fenner School of Environment and Society, Australian National University, Canberra ACT 0200, Australia

The Yangtze is the largest river basin in China and home to over 400 million people. In recent history, and especially during 1950s1970s, extensive lakes and oodplains were reclaimed as polders for agriculture and rural development. Consequently, the ood retention capacity was decreased, many lakes were disconnected from the main channel of the Yangtze by embankments and sluice gates, and eutrophication was common. It is anticipated that there will be a greater frequency of extreme oods and droughts in the basin according to climate change scenarios. WWF commenced a programme in 2002 in partnership with government agencies and local communities to reconnect three lakes (Zhangdu, Hong and Tian-e-zhou) in Hubei Province to the river by opening sluice gates seasonally and improving lake management. The resilience of the lake environment to climate change and the livelihoods of local people were enhanced. The measures assessed here highlight: (a) the need for adaptation programmes to concurrently improve livelihoods and reduce exposure to physical risks; (b) the need to build the capacity of people and institutions; and (c) the value of decentralized adaptation as compared with new infrastructure investments. Keywords: China; climate adaptation; climate change; freshwater management; Yangtze River

1. Introduction
The Yangtze River is the longest river in China and Asia and the third longest in the world. Running 6,300 km from the Tibetan Plateau to the East China Sea, the river system drains an area of 1,800,000 km2 in 19 provinces of China, and over 400 million people live in the basin (Figure 1). Its average annual runoff is about 996 billion m3 (BCM), accounting for about 36.5% of Chinas freshwater resource. The Yangtze River Basin accounts for 40% of Chinas gross domestic product (Yang et al., 2009).

The Yangtze River supports diverse ora and fauna that are well adapted to constantly changing water levels and ow; it has some of the richest biodiversity in Asia with over 340 species of sh alone. Consequently, WWF designated the basin as a Global 200 site and committed to conserve it from the mountain to the sea (WWF, 2007). Until late last century, the river and its lakes formed a complex wetland network fullling important ecosystem functions such as serving as the spawning and feeding grounds for sh and retaining summer oodwaters. However,

B *Corresponding author. E-mail: jianglg@igsnrr.ac.cn


CLIMATE AND DEVELOPMENT 1 (2009) 241248 doi:10.3763/cdev.2009.0023 # 2009 Earthscan ISSN: 1756-5529 (print), 1756-5537 (online) www.earthscanjournals.com

242 Yu et al.

FIGURE 1 Location of the Yangtze River basin; the central Yangtze is shown in the dashed frame

dam and dyke construction in Hubei Province, which took place on an especially large scale in the 1950s 1970s, has cut off 1,066 lakes covering over 2,000 km2 from the Yangtze main stem. Most of the lake shore area was converted to polders, where agriculture was developed. Altogether, the total wetlands area was reduced by 80% and ood water retention capacity declined by 75%, or 2.8 billion m3. The fragmentation of the river lake wetlands complex caused the interruption of hydrological, ecological and geochemical processes. Four major oods between 1991 and 1998 resulted in thousands of deaths and billions of dollars of direct economic losses. The lack of hydrological connection between lakes and the Yangtze River blocked the seasonal migration of sh and freshwater cetaceans and reduced the purication functions of wetlands and lakes. Disconnected lakes have become highly polluted from agricultural return ows, industrial and domestic discharge, and aquaculture. Recently, higher air and water temperatures associated with global climate change have exacerbated eutrophication and further reduced water quality.

Detecting impacts from climate change is statistically challenging, and circulation models of projected impacts from climate change do not have sufcient certainty or resolution to determine specic shifts of eco-hydrological regime in the central Yangtze region. Nevertheless, the Intergovernmental Panel on Climate Change reports suggest that the frequency and severity of extreme weather events is increasing and will continue to increase in coming decades even if mean annual precipitation may not shift signicantly (Bates et al., 2008). Recent extreme weather events have had powerful negative effects on the hydrology of the riverlakes complex over relatively short periods of time (weeks or months), but these effects have been sufciently strong to have altered sensitive species and ecosystems, as evidenced by oods in the 1990s, droughts in 2006 in Chongqing Municipality and in 2007 in the Dongting and Poyang lake areas, and the severe snowstorm in the central and lower Yangtze basin in January 2008. This paper describes climate change adaptation initiatives in the central Yangtze River by WWF and its partners. This assessment was undertaken

CLIMATE AND DEVELOPMENT

Freshwater management experiences from the central Yangtze in China 243

using an analytical framework as one of six case studies for a larger assessment of autonomous adaptation, as reported by Pittock (2009). Here, we provide an overview of the challenges facing this portion of the basin, the plan and implementation process, and the lessons learnt that may be relevant to other freshwater climate adaptation projects in the Yangtze, the rest of China, and beyond.

At 54 stations, annual average precipitation showed a downward trend, with a signicant decrease at eight (Jiang et al., 2007).

3. Projected climate trends


Climate scenarios for the middle reaches of the Yangtze River basin were simulated based on the ECHAM5/MPI-OM model by Jiang et al. (2007) in three IPCC emissions scenarios for the period of 2001 2050: SRES-A2 (high emissions), SRESA1B (moderate emissions) and SRES-B1 (low emissions). Under all three emissions scenarios, in the middle and lower reaches of the Yangtze River, annual precipitation does not change signicantly between 2001 and 2050, but the rate of inter-annual variability increases. This might cause more extreme climate events in the Yangtze Basin (Editorial Committee, 2007). These simulation results should not be taken as denitive. Other models are likely to produce different results. Moreover, air temperature trends are generally treated with much higher condence than precipitation or evapotranspiration trends. Yet it is reasonable to conclude that measures to reduce the impacts of oods and droughts may substantially aid adaptation to climate change in the central Yangtze basin.

2. Recent climate trends


Air temperature is frequently reported in climate data for freshwater ecosystems because it often represents the best quality and most spatially comprehensive data. In the Yangtze River basin, the average annual air temperature ranges from 188C in the south to 148C in the north, with most regions experiencing temperatures between 16 and 188C. Mean air temperature in the Yangtze basin has risen beyond that of the period 19611990: in the period 19912005 it rose 0.468C and for the period 20012005 it rose by 0.718C (Jiang and King, 2004; Jiang et al., 2007). Precipitation is another important climate variable inuencing river inows. The IPCC suggests that the timing, amount, form (rain vs. snow) and intensity of precipitation is shifting in many regions (Bates et al., 2008). The middle reaches of the Yangtze are deeply inuenced by a monsoon climate. Precipitation mainly occurs in summer time. Variation within and between years in the amount of precipitation is substantial for this region. Historically, severe ood disasters follow continuous rain events that span the whole basin. Between 1960 and 2005, mean annual precipitation for the whole of the Yangtze River basin was 1126.7 mm. Due to circulation variation and the impact of topography, temporal and spatial distribution of rainfall is very uneven, ranging from a low of 600 mm to a high of 1600 mm in the Poyang Lake (Jiang et al., 2007). Precipitation patterns are shifting with climate change over a large region of the basin. Data from 147 stations during 1960 2005 showed that the average annual rainfall increased at 93 stations, with a signicant increase at 19 stations.

4. Designing appropriate responses


In our initial assessment, we hypothesized that the climate resilience of the Yangtze River lake complex had been signicantly weakened as a result of the disconnection of the lakes from the river. In the past these connections were typically seasonal, with high connectivity during the monsoonal oods and low connectivity during the winter. We postulated that the climate adaptation capacity of these ecosystems and the livelihoods of people dependent on their ecosystem services could be improved by restoring the seasonal ecohydrological connections between the lakes and the Yangtze River. In 2002 WWF commenced a programme to reconnect lakes in Hubei province to the

CLIMATE AND DEVELOPMENT

244 Yu et al.

Yangtze River by opening the embankment sluice gates to facilitate more sustainable lake management. The programme focused on three lakes: Zhangdu (40 km2), Hong (348 km2) and Tian-e-zhou (20 km2). However, restoring connectivity alone was insufcient without addressing the other human-induced threats to ecosystem health. We also postulated that helping residents dependent on freshwater ecosystem services develop more sustainable livelihoods was a necessary means of ensuring the long-term climate-adaptive capacity of the ecosystems, particularly when average income is just USD1.34 per day (Li et al., 2005). In conjunction with this work, WWF formed partnerships with government agencies and others to explore potential solutions and to develop a privatepublic consensus for more sustainable river basin management. WWF worked to facilitate the adoption of mechanisms and processes that contribute to a longterm and sustainable approach to manage ood risk and conserve wetlands, improve livelihoods and stem the massive loss of biodiversity in the central Yangtze. In the context of climate change, these types of actions represent autonomous adaptation (Bates et al., 2008). The programme chose several overlapping goals: promoting community-based and wise use of wetlands, re-linking riverlake connections, and integrating the management network of protected areas of wetlands at both basin and national levels. To achieve them, the programme employed four basic approaches: establishing demonstration projects to develop best practices, direct policy advocacy, public education initiatives for resource users and public schools, and building cooperative networks between policymakers, resource managers and communities. In particular, it is believed that success would be more likely if demonstration sites were established with the close collaboration on-site of local governments and communities.

Chinas Wuhan Project Ofce, project partners (especially local government) or contained in an internal programme assessment report (DCruz and Yu, 2006). Starting from the summer of 2004, the sluice gates at the Tian-e-zhou, Zhangdu and Hong lakes have been reopened seasonally, with removal or modication of related illegal and uneconomical aquaculture facilities in the lakes. The reconnection restored the natural seasonal ooding and enhanced the wetlands capacity for water purication and ood retention. In 2005 in Zhangdu Lake, 285.6 106 m3 of ood waters from the Yangtze were safely stored. Further, we estimate that more than 5.26 million juvenile sh of 14 species were able to enter the lake as part of their seasonal migration. Consequently shery production in 2005 increased by more than 17%. The success of these changes led to the Anhui Provincial Government opening the sluice gates at the Baidang lake (40 km2) from 2006. Altogether, the approximately 448 km2 of wetlands now reconnected to the Yangtze River can store up to 285 million m3 of oodwaters, reducing vulnerability in downstream areas, although this has not yet been tested in practice. Cessation of unsustainable aquaculture, better agricultural practices and reconnection to the Yangtze River has reduced pollution levels in these lakes and improved water quality. Pollution levels fell at Hong Lake from national pollution level IV (t for agricultural use only) to II (drinkable) on Chinas ve-point scale. Subsequently, the Anhui Provincial Government has reconnected a further eight lakes at Anqing covering 350 km2.

5.1. Policy outcomes


Re-linking the disconnected lakes to the mainstream of the Yangtze has been included in the work plan of the local governments in Hubei and Anhui provinces. The practice has also been recognized and promoted by the central government through the Action Guideline of China Hydro-biological Resources Cultivation and

5. Programme results and outcomes


Unless otherwise stated, the data cited below is based on unpublished project data from WWF

CLIMATE AND DEVELOPMENT

Freshwater management experiences from the central Yangtze in China 245

Protection Initiative issued by the China State Council on February 14, 2006. With the support of this programme, WWF with other partner organizations submitted a policy recommendation report to the State Council on promoting integrated river basin management (IRBM) in China in 2004, and many of their recommendations have been accepted and implemented across the country (CCICED, 2004). The biennial Yangtze Forum was established in 2005 and has met three times so far to bring together key government departments and other stakeholders to share perspectives, develop sustainable economic policies, integrate data, promote adaptation to climate change and develop a vision for harmonious management of the entire river.

national government for ve years. The compensation is at least the same or even higher than their income from previous farming practice.

5.3. Environmental outcomes


The programme has restored and protected a total of over 400 km2 wetland. At Zhangdu Lake, 60 km2 of lake and marshland were designated as a nature reserve by the Wuhan Municipal Government. The Hubei Provincial Government approved a master plan for wetland conservation in 2006, which committed to protect an additional 4,500 km2 of wetlands by December 2010. In the Anhui province, in total 800 km2 of lake beneted from reconnection to the river. To strengthen the effectiveness of wetland conservation efforts in the Yangtze River basin, a Wetland Conservation Network was established in 2007, which links managers of 17 nature reserves (12 recently designated) covering 4,500 km2. Climate adaptation measures are now being promoted by this network. The ecological condition of the wetlands has greatly improved. Taking Hong Lake as an example, the aquatic vegetation has been restored and many water bird species have returned to the lake, including the endangered oriental white stork that abandoned the lake 11 years before. Water quality in the demonstration sites improved from national Class IV to II (i.e. suitable for drinking after simple processing). Furthermore, over 80% of xed shing nets were removed, which improved habitat quality and connectivity. The provincial government has pledged to remove all shing nets within two years from the site. Wildlife diversity and the population of many species have increased. Twelve migratory sh species returned to the lakes. Hong Lake supported only 100 herons and egrets when polluted, but after restoration 45,000 wintering water birds and 20,000 breeding birds returned. Tian-e-zhou Lake is the site of the managed popu` lations of the threatened Pere Davids Deer and the Yangtze nless porpoise, with the latters

5.2. Livelihood outcomes


Opening the sluice gate to re-link the lakes with the Yangtze River restored the seasonal migration of sh and introduced wild sh fry from the river, which resulted in the increased catch. For example, opening the sluice gate at the Zhangdu Lake in June 2005 introduced about 5.26 million fry. Six months later, the catch increased by 17.33% compared to that in the previous year. Similarly, the catch increased by 15% in the Baidang Lake. The programme helped local farmers to develop sustainable aquaculture by introducing high economic value sh species and reducing sh feeding inputs. As a result, income from aquaculture increased by 30%. The programme also facilitated the development of certied eco-sh farming of 412 households in Hong Lake, whose income from shery increased by 2030% on average. To reduce agricultural pollution, the programme also supported bamboo eco-farming among communities around the Zhangdu Lake. Bamboo farming has not yet increased the local farmers income directly, because harvesting will only commence in 2009. However, bamboo farming is listed in the national Grain for Green Project (a programme to return steep croplands to forest); farmers receive compensation from the

CLIMATE AND DEVELOPMENT

246 Yu et al.

population growing from 24 to 40. Surprisingly, the ice sh (Pseudolaubuca engraulis) reappeared after a 20-year absence.

6. Discussion 6.1. Factors for success


Local communities and municipal and provincial governments were motivated by better access to high-quality water, diversied local economies, increased incomes and an improved environment. The national government agencies were also motivated by the need to reduce ood risks, although the use of the reconnected lakes to store oodwaters is not supported by all local governments. The programme appears to have inspired widespread restoration and protection of many wetland sites by the relevant government authorities. The adoption of the new wetland policies and regulations by the relevant government agencies aids the sustainability of these measures. For example, Hubei Provincial Government resolved to intensify efforts to protect Hong Lake by allocating RMB73 million for its restoration and protection during 2006 2010. Local governments have issued ofcial documents to maintain seasonal connections between the lakes and Yangtze River at several lakes. The operation of the sluice gates in Tian-e-zhou oxbow and Zhangdu Lake have been modied to support re-linkage efforts and the alternative livelihood activities. Technical expertise has been engaged to help further diversify the alternative livelihood activities, such as through enhanced fruit production. The skills training and the increased income deriving from the new livelihood measures provided strong incentives for farmers to participate. However, these successes have not come easily. Altering ood control measures is controversial in any society. Demonstrating that adaptations can work in the eld was vital to learn by doing and to secure external support for wider application at provincial and national scales. Attentiveness to the needs of governments and

other stakeholders was essential for gaining support and ownership. In this context the motivation for local actors to participate in the programme came from the chance to improve livelihoods (Schipper, 2007), while national and provincial agencies were motivated by the potential to reduce the physical risks of ooding, pollution and biodiversity loss (Adger et al., 2005). We attribute the programmes success to:
B

Partnerships: Establishing and building strategic partnerships with key organizations (including donors) ensured the successful delivery of the goal and targets. It was equally, if not more important, to build capacity within partners to sustain these successes. Demonstrations: Demonstration sites allowed for attention to be focused on key issues and to seek practical and commonly agreed solutions, using new and innovative approaches; they also served as a valuable communications tool to advance the goal of the programme. Flexibility: The ability of the programme team to adapt to the needs of the stakeholders in the face of changing policies and priorities was a critical factor in ensuring that the outcomes were useful and sustainable. Learning: The proactive learning by doing approach adopted by the programme contributed to organizational learning within WWF and within the key partner organizations. Facilitation: It was important that the programme positioned itself strategically in order to be able to provide a common platform to facilitate discussions between stakeholders and beneciaries to develop consensus-based solutions. Communication: It was important to recognize the role of effective outreach to target audiences to achieve the goals of the programme and to sustain the outcomes. Mainstreaming: It was important to ensure that the programmes objectives and interventions were linked to the governments policies and priorities. This allowed the outcomes and lessons learned to be used to improve policy

CLIMATE AND DEVELOPMENT

Freshwater management experiences from the central Yangtze in China 247

and practice and ensure support for the programmes goals. These success factors appear consistent with the systematic social learning promoted by Lee (1993) for more sustainable environmental management. Further, these interventions emphasize the importance of enhancing the capacities of people and institutions in undertaking effective adaptation.

6.2. Magnication
There are hundreds of sluice gates along the Yangtze River that cut off lakes, so there is considerable potential to scale up this approach. This approach to Yangtze oodplain restoration represents a step towards better operation of existing infrastructure rather than further engineering-led interventions. It is substantially decentralized and largely applies existing knowledge. Further lessons and recommendations based on this programme have been proposed to national agencies in the Yangtze Conservation and Development Report 2007 (Yang et al., 2009). Further, this oodplain restoration strategy offers an alternative to the maladaptation of cutting more wetlands off from the river, as is proposed at Poyang Lake to manage droughts and oods (Water Resources Department, 2008). All the problems discussed in this paper for the demonstration region are also problems for the broader middle and lower reaches of the Yangtze, including the Poyang and Dongting lakes regions. At the Yangtze basin level, the programme has provided case studies that are inuencing the policies of key institutions in river basin management and climate change adaptation, such as the biennial Yangtze Forum. The Changjiang (Yangtze) Water Resources Commission has initiated a basin master-planning process, which is drawing upon key lessons from the programme to incorporate climate change adaptation and wetlands conservation measures. At the national level, the programme promoted the adoption of re-linking river and lakes as a key

measure in related policies. These efforts have inuenced other related policy changes at national level, such as the National Wetland Conservation Project, the draft National Wetland Conservation Regulations, the proposed National Natural Reserve Management Law and the national river basin planning process that is currently under way. The programmes many interventions for more adaptive river and lake management at the site, county, provincial, basin and national scales have been mutually reinforcing; they have facilitated more effective and efcient management and ensured equitable outcomes at both the site level and the legitimacy of the actions consistent with Adger et al. (2005). Future challenges include improving cross-sector policy development and enhancing accountability for the implementation of agreed-upon policies.

7. Conclusions
This case shows that restoring freshwater ecosystems has increased the resilience of the environment and economy to extreme weather events and eutrophication, which are predicted to be exacerbated with climate change. The connectivity between the Yangtze River and its oodplain wetlands and lakes has been reinstated. Assisting local sh farmers to adopt more sustainable practices has enhanced their livelihoods and the environment, and made these communities less vulnerable to extreme events. Working in partnership with government agencies has ensured that these new practices are now mainstreamed into related policies, plans, nancing and routine operations, and facilitated their adoption in other provinces. These shifts in behaviour highlight the need to concurrently improve livelihoods and reduce exposure to physical disasters that may be exacerbated by anthropogenic climate change, the need to build the capacity of people and institutions to manage changes in behaviour, and the opportunities for decentralized adaptation that does not rely on new infrastructure investments.

CLIMATE AND DEVELOPMENT

248 Yu et al.

Acknowledgements
The WWF-HSBC Yangtze Programme was funded by HSBC and managed by WWF-UK as part of their 2002 2006 Investing in Nature Partnership. The WWF-HSBC Climate Partnership funded this research. This research draws on the eldwork of a considerable number of WWF staff, partners and donors, who cannot all be named but whose contributions are greatly appreciated.

References
Adger, W. N., Arnell, N. W. and Tompkins, E. L., 2005. Successful adaptation to climate change across scales. Global Environmental Change (A), 15(2). 7786. Bates, B. C., Kundzewicz, Z. W., Wu, S. and Palutikof, J. P., 2008. Climate Change and Water. Technical Paper of the Intergovernmental Panel on Climate Change. IPCC Secretariat, Geneva. CCICED (China Council for International Cooperation on Environment and Development), 2004. Promoting Integrated River Basin Management for Living Rivers in China. www.cciced.org/2008-03/04/content_ 11509931.htm. DCruz, R. and Yu, X., 2006. Final Evaluation Report of WWF-HSBC Programme of Restoring the Web of Life in the Central Yangtze, China. WWF, Beijing, China.

Editorial Committee, 2007. Chinas National Assessment Report on Climate Change. Science Press, Beijing, China. Jiang, T. and King, H., 2004. Climate Change and Yangtze Floods. Science Press, Beijing, China. Jiang, T., Su, B. and Hartmann, H., 2007. Temporal and spatial trends of precipitation and river low in the Yangtze River Basin, 1961 2000. Geomorphology, 85. 143 154. Lee, K. N., 1993. Compass and Gyroscope: Integrating Science and Politics for the Environment. Island Press, Washington, DC and Covelo, CA. Li, Q., Yu, X. and Li, J., 2005. Analysis on wetland use change in Zhangdu Watershed, Hubei Province. Resources and Environment in the Yangtze Basin, 14(5). 600 604. Pittock, J., 2009. Lessons for climate change adaptation from better management of rivers. Climate and Development, 1(3). 194 211. doi:10.3763/cdev.2009.0021. Schipper, E. L. F., 2007. Climate Change Adaptation and Development: Exploring the Linkages. Tyndall Centre Working Paper No. 107. Tyndall Centre for Climate Change Research, Norwich, UK. Water Resources Department, 2008. Poyang Lake Project. Water Resources Department of Jiangxi Province, China. www.jxsl.gov.cn/list.jsp?classid963. WWF, 2007. Yangtze River and Lakes A Global Ecoregion. www.panda.org/what_we_do/where_we_work/ecoregions/yangtze_river_lakes.cfm. Yang, G., Ma, C. and Chang, S., 2009. Yangtze Conservation and Development Report 2009. Changjiang Press, Wuhan, China.

CLIMATE AND DEVELOPMENT

case study

Integrated river basin management in the Conchos River basin, Mexico: A case study of freshwater climate change adaptation
J. EUGENIO BARRIOS1, *, J. ALFREDO RODRIGUEZ-PINEDA2 and MAURICIO DE LA MAZA BENIGNOS3
1 2

WWF-Mexico, River Basin Management Program, Avenue Mexico 51, Col Hipodromo-Condesa, Mexico DF, C.P. 06100, Mexico WWF-Mexico, River Basin Management Program, C. Coronado 1005, Col Centro, Chihuahua, Chih., C.P. 31000, Mexico 3 WWF-Chihuahuan Desert Program, C. Coronado 1005, Col Centro, Chihuahua, Chih., C.P. 31000, Mexico

In Mexico, due to reduced and unevenly distributed hydrological resources and incipient water management capabilities, climate change adaptation in the water sector is recognized as an urgent issue. To derive lessons for climate change adaptation, this paper evaluates the results gained after ve years of an integrated river basin management (IRBM) programme in the Conchos River in northern Mexico. Autonomous adaptation measures assessed include: modernization of irrigation practices; pilot sustainable watershed management projects in the upper basin; development of an environmental ow assessment and a proposal to improve water allocation; and the creation of the Inter-institutional Working Group as a basin organization. These measures have improved river basin management, yet adverse outcomes were also observed, such as impacts of surface water efciency measures that were not managed in conjunction with groundwater. Key adaptation lessons derived include: the importance of multi-stakeholder participation in designing and implementing adaptive management measures; the need for signicant investment in transfer of expertise and capacity building; and the positive effect of linking local, national and international institutions. These results highlight the need for more investment in soft adaptive management in place of infrastructure. In the Rio Conchos, if these no regrets adaptation measures are consolidated in the following years, they will serve as a foundation to develop planned and more effective climate change adaptation programmes, and enhance institutional, environmental and societal resilience. Keywords: adaptation; basin; climate change; integrated river basin management; river; water

1. Introduction
Beginning in 2004, the alliance between WWF and the Gonzalo Rio Arronte Foundation (WWF-FGRA) commenced an integrated river basin management (IRBM) strategy in three Mexican river basins: the Conchos River in the state of Chihuahua, the Copalita-ZimatanHuatulco Rivers in the state of Oaxaca, and the San Pedro-Mezquital River in the states of Durango and Nayarit. The programme aims to improve water management in order to maximize the economic and social benets derived from
B *Corresponding author. E-mail: ebarrios@wwfmex.org
CLIMATE AND DEVELOPMENT 1 (2009) 249260

water resources in an equitable manner, while at the same time preserving and restoring freshwater ecosystems. Water scarcity already impacts on large portions of Mexico, especially in the arid north where annual water availability per person is only 1,750 m3 compared to the national mean of 4,416 m3/person/year (CONAGUA (Comision Nacional del Agua), 2007). This unbalanced hydrological distribution jeopardizes the social and economic development of approximately 60% of Mexican territory and 77% of the Mexican population.

doi:10.3763/cdev.2009.0024 # 2009 Earthscan ISSN: 1756-5529 (print), 1756-5537 (online) www.earthscanjournals.com

guez-Pineda and De la Maza 250 Barrios, Rodr

Climate change forecasts predict that there will be further impacts on water resources in Mexico. For example, less runoff and more frequent droughts are expected in the arid north. Furthermore, as in other parts of the world, demand for consumptive water is expected to exceed supply, partly as a result of the impacts of global warming on water supplies up to 2025 (Vorosmarty et al., 2000). Consequently, Mexico needs to develop effective climate change adaptations, particularly to cope with the exacerbation of water scarcity. This paper assesses the water sector adaptation implemented in the Rio Conchos basin. The authors consider that this sort of autonomous adaptation in the water sector (Kundzewicz et al., 2007; Bates et al., 2008) holds lessons that can aid the design of more effective climate change adaptation programmes. This case study was prepared as part of a larger review of freshwater autonomous adaptation projects by WWF, which makes up this volume. The Conchos River basin is one of the most important basins in northern Mexico, draining an area of 67,000 km2. It comprises 14% of the binational (Mexico USA) Rio Bravo/Rio Grande1 basin, which has a catchment area in Mexico of 226,280 km2 and annual runoff of 6,177 million m3/year; one-third of this volume (2,553 million m3/year) is provided by the Conchos River tributary. The Conchos River originates in the Sierra Madre Occidental, locally known as Sierra Tarahumara, at a mean altitude of 2,300 m above sea level, and makes a conuence with the Rio Grande/Bravo at the US Mexican border (Figure 1). The river and its tributaries are regulated by seven main reservoirs with a total capacity of 3,654 million m3. Annual precipitation in the basin ranges from 700 mm in the upper portion to 250 mm in the lower portion. Precipitation predictions for the next 100 years indicate a slight increment in rainfall in the study area (Hadley Centre, 2005), although this increase will be insignicant due to the increase in air temperature and evaporation,

and soil water decit increments (Raynalor and Rodrguez-Pineda, 2008). In Villasen addition to the low precipitation values, the basin is prone to long periods of drought, such as the most recent, which extended from 1993 to 2005 oz (Mun et al., 2006; Reyes-Gomez et al., 2006).

2. Water availability and use


Historical basin runoff is 2,244.7 million m3; 44% of this volume is produced in the upper basin (985.7 million m3) (DOF, 2008), the Sierra Tarahumara, and feeds the Boquilla dam the main source of water for the Delicias Irrigation District oz, 2007). and the Conchos River basin (Mun Water scarcity is the main threat to life in the Chihuahua Desert. Since the 19th century, the construction of irrigation projects and dams for agriculture has reduced hydrological variability; however, these adaptations were challenged by the 1993 2005 drought, which reduced river inows to 25% of the historical mean over the last 60 years. In 1999, the Rio Grande did not reach the Gulf of Mexico for the rst time in recorded history. In 2004, when the longest measured drought broke (Reyes-Gomez et al., 2006), the Conchos River surface water availability reached a normal value of 1,679 million m3 allocated to consumptive uses that was distributed as follows: 1,648 million m3 (98%) for irrigated agriculture, 23 million m3 (1.4%) for domestic use, 3 million m3 (0.2%) for industry and 5 million m3 (0.4%) for other uses. For non-consumptive uses, a volume of 2,311 million m3 was used for hydroelectric facilities. These values can be compared to those of 1995, at the beginning of the drought, when agricultural water use decreased to only 315 million m3 (CONAGUA, 2007, 2008a), and also to the additional 432 million m3 that Mexico is obliged to deliver to the USA from the Rio Conchos under the 1944 water treaty (Water Treaty, 1944).

CLIMATE AND DEVELOPMENT

Integrated river basin management in the Conchos River basin, Mexico 251

FIGURE 1 The Conchos River basin area, Mexico

CLIMATE AND DEVELOPMENT

guez-Pineda and De la Maza 252 Barrios, Rodr

The basin consists of three irrigation districts. The DR-005 Delicias covers 81% of the total irrigated area (i.e. 88,000 ha and 12,000 farmers). Likewise, it is allocated 83% of the basins total water volume (1,132 million m3/yr) in surface water rights. It represents by far the greatest impact on water extraction in the Conchos River (CONAGUA, 1997). The Jimenez Camargo, Camargo Delicias, Meoqui Delicias and the Aldama San Diego aquifers are the main groundwater sources for extensive agricultural areas and for urban uses. Of these, the Jimenez Camargo and Meoqui Delicias are overexploited at respective rates of 1.50 and 1.56 times the estimated natural recharge (CONAGUA, 2007). These hydrogeological basins, located under the main arteries of the Conchos, Florido and San Pedro Rivers, need to be managed to guarantee base water ows and also as future underground water storage to avoid greater evaporative losses due to climate change. Based on the air temperature increments forecast by the Hadley Centre (2005) and the IPCC or (2007), Raynal-Villasen and Rodrguez-Pineda (2008) developed evaporation and moisture decit scenarios. Results show that potential evaporation will increase by 2.0% to 7.3%, while the moisture decit will increase by 2.4% to 11.3% as air temperature increases by 1 8C to 3 8C over the next 30 and 100 years, respectively.

Recovering freshwater ecosystem functions through an environmental ow allocation, which is water that is needed for the environment. B Enhanced river basin governance through river basin councils and public participation. B Small-scale replicable projects to demonstrate rational water usage and natural resources management that can provide specic support to enhance living conditions in rural communities. B Public outreach and educational activities aimed at creating awareness of water scarcity, water values, and the role of water as a part of the environment, as well as communicating lessons learned from the project.
B

After ve years of implementation, this paper reviews the factors that aided or hindered the WWF-FGRA programme so that lessons from past and current water management experiences may serve as a foundation upon which to develop more reliable climate change adaptation policies and practices. As stated by several authors, water management involves adaptation, and there is a lot to be learned from autonomous adaptation (as undertaken in the Conchos River) for enhancing adaptation to climate change (Kundzewicz et al., 2007; Bates et al., 2008).

4. Results
The outcomes from the four major WWF instigated interventions that form the IRBM model in the Rio Conchos from 2004 2009 are described below.

3. Methods
WWFs intervention in the Conchos River is part of the Chihuahuan Desert Conservation Plan and the WWF Mexico freshwater conservation strategy. It is focused on implementing an IRBM model to conserve freshwater ecosystems; working with governments, the private sector, local communities, and non-governmental organizations; and implementing demonstration projects for the sustainable use of water. The project is being implemented through four main strategies:

4.1. Environmental ows as an IRBM adaptive management tool


The current water management paradigm in Mexico has promoted total extraction of water for consumptive uses. The proposal in the Rio Conchos to set up an environmental ow is focused on dening a new paradigm in which

CLIMATE AND DEVELOPMENT

Integrated river basin management in the Conchos River basin, Mexico 253

water needed for the environment both temporally and spatially becomes a limit on water extraction in order to foster long-term sustainability. Furthermore, this type of ecosystem-based approach was anticipated to increase the resilience of the environment and the local society to the impacts of climate change. The integration of environmental ows under scarcity conditions in arid regions, such as the Rio Conchos basin, raises many scientic, administrative, engineering and ideological challenges. First, (a) a scientic-based analysis is required to determine the quantity and quality of water that the river needs, as well as when it is required, in order to maintain ecosystem functions; (b) water administration must be modied to reallocate the water rights of current users; (c) hydraulic infrastructure must be operated under new rules to resemble natural ows; and (d) water users must understand that a owing river is not a waste of water but rather part of a healthy water basin. This is a new paradigm for water management that, although recognized in the Mexican National Water Law of 1992, has not yet been implemented. The Environmental Flow Assessment (EFA) for the Conchos River was a useful process to propose an ecosystem approach to water management for key stakeholders. It has also provided a mechanism to integrate previously fragmented and poorly accessible information. For instance, hydrological information was recovered from the Mexican National Water Agencys (CONAGUA) archives to form a comprehensive hydrological database. WWF, along with partners including the University of Texas, used this database to develop the water demand and supply model for the Conchos River using the Water Evaluation and Planning System (WEAP) (Patino-Gomez et al., 2007, 2008). This model is a key tool for developing water management scenarios based on the predicted inow variations that occurred during the last drought, and for the prediction of further variations due to climate change. An unexpected benet arising from this improved information system is its application

to ood management. Melchor Lopez (pers. commun., 2007), a specialist from CONAGUA in the Chihuahua Ofce, states that the EFA hydrological and climate database was an essential tool used to safely operate Las Virgenes dam when it was subject to an extreme precipitation event in 2007. As a part of the EFA process, ecological analyses were made based on habitat assessments and biological indicators (invertebrates, macroinvertebrates and riparian vegetation). The Historical Biological Index (HBI) and the Index of Biological Integrity (IBI) are now measured at 21 sites every two years, and are part of the monitoring system (Contreras-Balderas et al., 2005). All of this information was used by a group of experts to develop the EFA by applying the Building Block Methodology (BBM) (King et al., 2000). It is recognized as one of the most complete EFAs ever performed in Mexico, and it has been used as a model for other basins where WWF is working in Mexico and for the development of national standards and regulations. The EFA is being proposed as a strategy aimed at achieving the sustainable use of water in the basin by the year 2030. The strategy sets yearly goals to recover the currently overallocated 438 million m3 decit of water, as recogpez, nized by CONAGUA (pers. commun., Lo 2008). This water volume will represent a sustainable extraction in the basin for the recovery and conservation of river ecosystems, and will assist compliance with the water deliveries under the Rio Bravo/Grande international treaty (International Boundary and Water Commission/Comin sio Internacional de Lmites y Aguas IBWC/ CILA, 2009) from which one-third of the total water volume goes to the USA and two-thirds goes to the Mexican states in the lower part of the basin (Trueba and Goicochea, 2008).

4.2. Agricultural modernization


After the rst year of drought in 1994, the absence of rainfall resulted in conict between the farmers

CLIMATE AND DEVELOPMENT

guez-Pineda and De la Maza 254 Barrios, Rodr

and water authorities, who decided not to use the water stored in the main reservoirs since there was not enough water for all users. As a result, 8,000 ha of pecan trees were lost, as well as thousands of hectares of alfalfa and other perennial crops (Chavez, 2007). Thereafter, in the second and subsequent years, the Mexican farmers and water and agricultural authorities worked together in order to maintain or recover irrigated agriculture and to deliver the required amount of water to the Rio Bravo basin and the USA. Informed by a WWF assessment document (WWF, 2002), the government increased agricultural water efciency from 44 to 66% through the implementation of more efcient irrigation techniques in the Delicias irrigation district (CONAGUA, 2006, 2008b). Water demand management works were undertaken with an investment of US$140 million from the North American Development Bank, to assist Mexico to meet the downstream water deliveries required to full its treaty obligations with the USA. This investment was used to reduce irrigation water transmission losses by piping and lining earthen channels and by increasing the efciency of water application in the elds. In addition, a number of water licences were bought back on a voluntary basis and retired, thereby reducing irrigation water demand and increasing the reliability of the remaining allocations and the viability of the farmers concerned. Changes from low- to high-efciency irrigation practices and techniques have reduced the water volume per hectare for the main crops, as shown in Table 1. While good results were achieved, some problems arose, one of which was related to groundwater sources. Initially groundwater extraction was not capped and, as surface water allocations contracted, this resulted in the displacement of water extraction to the aquifers. Based on this experience, it has become clear to the farmers that both surface and groundwater must be jointly managed in order to reduce their vulnerability to water scarcity.

TABLE 1 Irrigation modernization Crop

efciencies

before

and

after

Irrigation volume (m3/ha) Before modernization After modernization 8 9 7.5 6.2 7.5

Volume reduction (%)

Pecans Alfalfa Onions Cotton Peppers

12.1 12.9 10.6 8.3 13.1

34 30 29 25 43

Source: Chavez, 2007.

In 2008 it was said that the local farmers and authorities had achieved success, because Mexico had complied with the international water treaty with the USA by delivering the required water to the Rio Bravo (IBWC/CILA, 2009). During this process, farmers had to drastically change how they used water in order to produce the same or an even greater quantity of crops using less water. Thus the modernization process has resulted in huge changes in the way farmers think about water, the introduction of modern techniques and changes in water resources administration aimed at a more equitable distribution of water among all users. Furthermore, the drought conditions have made people think about the role that forest and soil conservation plays in river basin management. As a result, farmers and state institutions are recognizing the importance of the upper basin, where the most signicant volume of water originates, and the key role of soil conservation and reforestation in the watershed.

4.3. River basin governance


The Conchos River Commission was created in 1999 as a part of the Rio Bravo Basin Council under federal water law. However, it has not been active, and its membership is legally restricted to water rights owners having a water

CLIMATE AND DEVELOPMENT

Integrated river basin management in the Conchos River basin, Mexico 255

concession, thereby limiting stakeholder participation. In 2004, WWF created an alternative organization called the Inter-institutional Working Group (GIT), which has become an independent forum for all stakeholders interested in the Rio Conchos, whether institutions, groups or individuals. The GIT was ofcially recognized by the Government of Chihuahua in 2005 when a collaborative inter-institutional agreement was signed with WWF. The GIT includes government representatives, water users, universities, indigenous communities, non-governmental organizations (NGOs) and representatives of economic sectors. The group has developed a river basin management plan that has been allocated nearly USD8 million in investments from federal and state government programmes as a result of collaboration between public institutions of the three levels of government (local, state and federal) in the watershed. Since the start of the programme, the GIT has carried out 65 activities with an investment of close to USD3.2 million in 2005, and 60 activities with an expenditure estimated at USD4.4 million in 2006. Currently, CONAGUA (2008a) is taking advantage of this process to promote the reactivation of the Rio Conchos River Basin Commission among water users; however the GIT, as a multistakeholder organization rather than an exclusive water user organization, is challenging the current scope of the river basin commissions dened by the National Water Law. The creation of the GIT as an informal institution has been one of the major successes in the IRBM strategy. Some of the key elements of this success are: (a) the promotion of the basin concept as a geographical area that depends on the same water source, which is not familiar to the average person, as shown by a survey in which 80% of the people in the main Rio Conchos cities does not know they were part of the basin (WWF, 2004); (b) a forum not owned by any one organization, at which government agencies can present, rene and gain support for their programmes; and (c) participation of a

broad range of stakeholders in developing a common vision for the basin and implementing projects to achieve this vision.

4.4. Demonstration projects to support local communities


Indigenous communities made up of Raramuris, Tepehuano, Pima and Guarijo groups live in the upper part of the basin, known as Sierra Tarahumara. The development of this region has been a challenge throughout Mexican history. WWF and its partners have focused on integrating the indigenous communities into the IRBM process and supporting them to improve their livelihoods through better water supply and sanitation, forest conservation and biodiversity conservation. Two examples of livelihood improvement illustrate the outcomes of this community-based work. The rst is the implementation of a soil conservation project in the Choguita Aguatos microbasin that has restored 1,273 ha of eroded terrain and has directly beneted 433 people by improving their lands, living conditions and access to water. The second pilot project was designed to complement the soil and forest conservation practices with livelihood modules to promote a sustainable rural water management programme in the Upper Conchos River basin. Modules include: (a) rain water capture, (b) vegetable or backyard gardens, (c) dry toilets based on ecosan technology,2 and (d) reuse of grey water and kitchen organic waste. In addition, actions to protect potable and spring water sources and to support community water committees are taking place. Currently, 26 backyard gardens, each of 100 m2, have been built to cover a total area of 2,600 m2, as a result of which around 200 people have beneted directly. During the summer 2007 harvesting season, each garden owner was able to produce between four and 12 species of vegetables and capture at least 10 m3 of water for the winter. Families have

CLIMATE AND DEVELOPMENT

guez-Pineda and De la Maza 256 Barrios, Rodr

decreased their vulnerability to water scarcity and low temperatures during the winter. Currently, this work is being strengthened by supporting community water committees. Furthermore, WWF and its partners are developing an ecological tourism project in Ejido Panalachi, and a pilot project on payment for environmental services in the Choguita Aguatos basin, in order to further develop a proposal to improve the longterm living conditions of the indigenous communities in the Sierra Tarahumara, who are the main owners of the forest watersheds.

5. Discussion
The severe drought in the Conchos River from 1993 to 2005 was an extraordinary opportunity to study different responses and actions taken by water users, authorities and NGOs. However, it should be taken into account that these responses and actions were not planned under a climate change vulnerability and adaptation framework, in which the criteria of effectiveness, efciency, equity and legitimacy, as proposed by Adger et al. (2004), were considered in measuring adaptation success. These actions were a conscious response to climate variability and water shortage. In this sense, they could become the foundation of further climate change adaptation measures (Kundzewicz et al., 2007; Bates et al., 2008). The IRBM programme for the Conchos River basin was planned with a number of aims in mind: as a process to promote a basin-wide view; to integrate actions to overcome water overextraction and conserve freshwater biodiversity; to develop better basin governance; and to improve livelihoods of indigenous communities in the headwaters. As a result of ve years work, there are valuable implementation experiences the no regrets autonomous adaptation measures from which to learn, in order to develop a more effective climate adaptation programme for the basin, as well as for the country.

The EFAs implementation strategy in the Rio Conchos has strengthened the scientic understanding of the basin from elds such as hydrology, ecology and social sciences. It has allowed the proposal of freshwater ecosystem water requirements to CONAGUA, water stakeholders and the society for the sustainable management of the river ecosystem as a common vision for all. Thus the benet of maintaining and restoring a socioecological river system that was not considered at the time of the 19932005 drought is now being contemplated. This vision promotes, in a practical way, the convergence of vulnerability and resilience understanding (Adger, 2006), and therefore a better framework to adapt to climate change. The North American Development Banks USD140 million investment to improve irrigation infrastructure is the most important adaptation action taken in recent years. It has resulted in savings of between 25 and 43% in crop water use, and an increase in irrigation efciency from 44 to 66%. However, the success of such actions depends on spatial and temporal scales and should not be addressed simply in terms of the objectives of individual adaptors (Adger et al., 2004). Notably, this investment was driven in large part by the different but converging interests of local, state, national and US authorities (through the Mexico USA water treaty) in managing water scarcity. To date, this action has achieved the objective of reducing agricultural water demand, but it has decreased the groundwater recharge that was taking place in the unlined irrigation channels before modernization and, consequently, it could be part of the increment in uoride and arsenic concentrations of natural origin (Rodrguez-Pineda et al., 2005), that affects groundwater drinking supplies in the middle basin by reduction of fresh water inltration and heavy groundwater withdrawal mainly for agricultural use (Mahklnecht et al., 2008). This illustrates the need for adaptation measures to be planned with a broad range of expertise and for stakeholders to avoid such adverse outcomes.

CLIMATE AND DEVELOPMENT

Integrated river basin management in the Conchos River basin, Mexico 257

The water demand reductions that have been achieved could be a temporary solution, since they are insufcient to sustain irrigation farming if available water is reduced by more than 25%, as occurred during the last drought. The economic return from agricultural production in the region depends primarily on more than 60,000 ha of alfalfa as a cash crop (CONAGUA, 2008a, 2008b) and on high-value crops such as pecans, onions and peppers. According to farmers interviewed in the Delicias district, adaptations that would enable even less water use while maintaining or increasing economic returns are hindered by a lack of capacity to diversify their production and cultivate high-value crops in domestic, regional and international markets. As such, a sustainable adaptation appears to rst require adjustments in policies, institutions and attitudes in order to establish enabling conditions, which then would facilitate technological and infrastructural changes (Schipper, 2007). As stated by Liverman (1999) in its adaptation studies on drought in Mexico, over the longer term, improvements in irrigation efciency have reduced drought vulnerability in some irrigation districts, and new proposals for decentralized management and water pricing may allow more exible adjustment to water supply variations. In this sense, it has become clear through the Rio Conchos experience that modernization of the agricultural sector in Mexico will not be enough to achieve an effective adaptation to reduce vulnerability. The GIT has played a key role as a river basin entity that promotes consideration of proposed measures from diverse perspectives, equity of outcomes and legitimacy of decision making. Until now, adaptation actions have been decided mainly by the authorities, with limited participation of stakeholders other than the farmers. It is clear that decisions and actions taken prior to the 2004 process would have been better in the post-2004 environment, since they would have been taken considering other water users and the river ecosystem. For example:

A USD140 million investment in the irrigation district compared with the USD1.4 million invested in water and sanitation services in the Sierra Tarahumara in 2004 (World Bank, 2007). B Water savings were not allocated to recover the river ecosystem, but were rather proposed to increase the irrigation surface area. Currently, there is a proposal to allocate these savings to the environment and restore river ecosystems. B There was no consideration of groundwater impacts due to the reduction of water inltration from unlined channels and their impact on drinking water sources. Now it is clearer that surface and groundwater must be managed together. B The Conchos River Commission was not active at all in promoting participation or a river basin view. Currently, the GIT is playing the role of a river basin entity.
B

Consequently the authors contend that GIT, as a multi-stakeholder, establishes a better basis for adaptation decisions than a single stakeholder view, which avoids adverse outcomes, increases the effectiveness of adopted measures and enhances resilience. However, the GIT is overly dependent on voluntary work, as funding for this basin-scale management institution remains uncertain. As in other places in Mexico, it is clear that infrastructure development has been preferred in the Conchos River basin over soft infrastructure actions, which are not yet considered a priority. As stated by Adger et al. (2004), the benecial effects of soft engineering approaches are uncertain; however, in the Conchos River, these types of measures have shown good results on investment, such as the case of the GIT. A key barrier to soft infrastructure interventions is the limited nancial support, if any, that they are receiving. This is especially important because the recently presented Mexico Special Program on Climate Change 2008 2012 (PEF,

CLIMATE AND DEVELOPMENT

guez-Pineda and De la Maza 258 Barrios, Rodr

2009) is placing hydraulic infrastructure at the core of the adaptation strategy. It will be a signicant programme, but its effectiveness depends upon the active and informed participation of stakeholders in its planning and implementation. The IPCC has also raised this concern in the discussion of demand-side vs. supply-side adaptation options, indicating that it is a matter of uncertainty vs. certainty in results, and proposing that although supply-side options have environmental consequences, they can be alleviated in many cases (Kundzewicz et al., 2007, pp. 197 198). Based on the Conchos River experience, demand-side adaptation must be considered as the rst option, since it would not allow further environmental degradation; so the development of good water governance to support demandside options is a key issue. At the local level in the Sierra Tarahumara, concrete experiences illustrate how traditional projects to improve human well-being are the best way to reduce indigenous communities vulnerability to climate variability. At this time, these projects are covering a small amount of the population, since actions to support local indigenous communities must respect their customs and traditional decision-making processes. This is a good example of how legitimacy adds to project implementation. However, through this experience it has also become clear that if these projects are not complemented by local organization and community participation in the river basin governance, indigenous communities could be threatened by water management decisions. As stated by Adger (2006), equity within the decision-making process is as important as the equity of the outcome in reducing vulnerability. Since the Conchos River is part of a larger transboundary basin, the challenge ahead is not only to develop an evidence-based adaptation strategy but also to ultimately integrate this strategy with the entire Rio Bravo basin and into the Mexico USA international water agreement. Integrating climate change adaptation across these geo-

political entities is a challenge, since it requires action from the local to the international level. While it provides one framework for action, the Mexico USA water treaty (IBWC/CILA, 2009) has not been signicantly revised since 1944, during which time climate change and shifting human demands have been constantly imposing new challenges. Consequently, amendments to the agreement to manage for climate change should be sought.

6. Conclusions
In the process of implementing the IRBM programme in the Conchos River, the Alliance WWF-FGRA has developed an effective model for adaptation in water and river basin management that combines an ecosystem-based approach, public participation and support of indigenous communities to improve their livelihoods based on biodiversity conservation. Adaptive management in the Rio Conchos basin has proven more of a political than a technical challenge, as it has elsewhere in the water sector (Allan, 2003). Stimulating political support for adaptive water management in the Rio Conchos basin has required extensive investment in awareness raising, transfer of expertise and other capacity building to overcome lack of information, weak institutions and limited public participation. Most of the useful experiences from this process arise from comparing the decisions taken both before and after the 2004 process. Previously, most decisions were made by each sector or agency in isolation, and some decisions represented maladaptation. Since 2004, the IRBM approach provides a comprehensive framework for considering a broad range of expertise and interests, and for decision making on adaptation measures that is more likely to be effectively implemented without adverse outcomes. In this case, the IRBM in the Conchos River is the most important and clear adaptation strategy.

CLIMATE AND DEVELOPMENT

Integrated river basin management in the Conchos River basin, Mexico 259

The Mexico Special Program on Climate Change 2008 2012 (PEF, 2009) favours hydraulic infrastructure projects. However, based on the Conchos River experience, adaptive water management should instead favour public participation, soft engineering approaches, an ecosystem-based approach and demand-side adaptation options prior to any further physical alteration of water sources.

Acknowledgements
This programme has been developed by the Alliance WWF Mexico and the Gonzalo Rio Arronte Foundation, with support from HSBC, USAID, WWF-UK and The Coca Cola Company. The main participants in this programme are the Government of the State of Chihuahua, the National Water Commission (CONAGUA), the National Forestry Commission (CONAFOR), the communities of the Ejido Panalachi and Choguita, and the NGOs Alcadeco and Profauna.

Notes
1. The Rio Bravo is known in the USA as the Rio Grande. 2. Ecological sanitation (ecosan) solves sanitation problems by recovering and reusing the resources contained in excreta and wastewater (www.ecosan.org).

References
Adger, W. N., 2006. Vulnerability. Global Environmental Change, 16(3). 268 281. Adger, W. N., Arnell, W. N. and Tompkins, L. E., 2004. Successful adaptation to climate change across scales. Global Environmental Change A, 15(2). 77 86 Allan, J. A., 2003. Integrated water resources management is more a political than a technical challenge. Developments in Water Science, 50. 9 23. Bates, B. C., Kundzewicz, Z. W., Wu, S. and Palutikof, J. P. (eds), 2008. Climate Change and Water. Technical Paper of the Intergovernmental Panel on Climate Change. IPCC Secretariat, Geneva. Chavez, H., 2007. Personal communication. Farmer leader of the Irrigation District 05 Delicias.

CONAGUA, 1997. Programa Hidraulico de Gran Vision del Estado de Chihuahua 1996 2020. Comision Nacional del Agua, Subdireccion General de Programacion. Gerencia Regional Norte, Gerencia Estatal Chihua hua, Mexico. CONAGUA, 2006. Sexto Informe de Gobierno del Presidente de los Estados Unidos Mexicanos Vicente Fox Quesada. Derechos Reservados. Presidencia de la Republica. 1 Sept 2006. CONAGUA, 2007. Estadsticas del agua en Mexico. Comi sion Nacional del Agua, Mexico. www.conagua.gob. mx/Publicaciones. CONAGUA, 2008a. Meeting notes. Funcionarios de la CONAGUA sobre el programa de modernizacion y tecnicacion del Distrito de Riego 005, Delicias, Chi huahua, Mexico. Comision Nacional del Agua, Gobierno del Estado de Chihuahua, Asociacion Civil de Usuarios y el Convenio para la ejecucion de las acciones de modernizacion de las obras de infraestructura hidroagrcola y tecnicacion del riego en el Distrito de Riego 005. Cd. Delicias, Chihuahua, Mexico. CONAGUA, 2008b. Comunicacion con personal tecnico. Direccion Local en Chihuahua y Ocina de Residencia de la Comision Nacional del Agua (CONAGUA) en el Distrito de Riego 005. Delicias, Chihuahua, Mexico. Contreras-Balderas, S., Lozano-Vilano, M. and Garca Ramrez, M. E., 2005. Indice Biologico de Integridad, Version Historica (IBIH). El Ro Conchos, Chihuahua, Mexico Bioconservacion, A.C. WWF Programa Desierto Chihuahuense, Universidad A. de Nuevo Leon, San Nicolas, N.L., Mexico. DOF, 2008. Diario Ocial de la Federacion Acuerdo por el que se da a conocer el resultado de los estudios de disponibilidad media anual de las aguas superciales en las cuencas hidrologicas Ro Bravo. 22 Sept 2008. www.ordenjuridico.gob.mx/Federal/PE/APF/APC/ SEMARNAT/Acuerdos/2008/22092008(2).pdf. Hadley Centre, 2005. Climate Change and the Greenhouse Effect. www.metofce.gov.uk/research/hadleycentre/pubs/brochures/2005/climate_greenhouse.pdf. IBWC/CILA, 2009. Tratado sobre Distribucion de aguas internacionales entre los Estados Unidos Mexicanos y los Estados Unidos de America. (Tratado de Aguas de 1944) 3 February 1944. www.sre.gob.mx/cila/. IPCC, 2007. Intergovernmental Panel on Climatic Change, Fourth Report. www.ipcc.ch/ipccreports/ index.htm. King, J. M., Tharme, R. E. and de Villiers, M. S., 2000. Environmental Flow Assessments for Rivers: Manual for the Building Block Technology. South African Water Research Commission Report No. TT 131/00.

CLIMATE AND DEVELOPMENT

guez-Pineda and De la Maza 260 Barrios, Rodr

Kundzewicz, Z. W., Mata, L. J., Arnell, N. W., Doll, P., Kabat, P., Jimenez, B., Miller, K. A., Oki, T., Sen, Z. and Shiklomanov, I.A., 2007. Freshwater resources and their management. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, M. L. Parry, J. P. Palutikof, P. J. van der Linden and C. E. Hanson (eds). Cambridge University Press, Cambridge, UK. 173210. Liverman, D. M., 1999. Vulnerability and adaptation to drought in Mexico. Natural Resources Journal, 39(1). 99 115. Lopez, Mario, 2008. Personal communication. CONAGUA, Technical Ofce Mexico. Lopez, Melchor, 2006. Personal communication. CONAGUA Chihuahua Technical Department. Mahlknecht, J. A., Horst, A., Hernandez-Limon, G. and Aravena, R., 2008. Groundwater geochemistry of the Chihuahua City region in the Rio Conchos Basin (northern Mexico) and implications for water resources management. Hydrological Processes, 22(24). 47364751. www.interscience.wiley.com. doi: 10.1002/hyp.7084. oz, C., Nun ez, D., Gadsden, H., Rodrguez, J. A., Mun Reyes-Gomez, V. and Hinojosa, O., 2006. Creation d un observatoire de la secheresse dans LEtat de Chihuahua au Mexique. Science et Changements Pla netaires Secheresse, 17(4). 467474. oz, E. I., 2007. Estudio de disponibilidad media anual Mun de agua en la Cuenca del Ro Conchos. Informe tecnico, Convenio KH12. WWF-Programa Desierto Chihua huense, Chihuahua, Mexico. o-Gomez, C., McKinney, D. and Maidment, D. R., Patin 2007. Sharing water resources data in the binational Rio Grande/Bravo Basin. Journal of Water Resources Planning and Management, 133(5). 416426. Patino-Gomez, C., Aparicio-Mijares, J., Sanvicente Sanchez, H. and McKinney, D., 2008. Water Management Information System in MexicoUnited States of America Border Region. IV International Symposium on Transboundary Waters Management, Thessaloniki, Greece, 1518 October. www.inweb.gr/twm4/abs/ PATINO%20GOMEZ%20Carlos.pdf. PEF (Poder Ejecutivo Federal), 2009. Programa Especial de Cambio Climatico 2008 2012 [Special Program on Climate Change 2008 2012]. Version consulta publica. www.semarnat.gob.mx/queessemarnat/co

nsultaspublicas/Pages/peccconsultacomplementaria. aspx. or, J. A. and Rodrguez-Pineda, J. A., Raynal-Villasen 2008. Posibles escenarios del impacto del cambio cli matico en la cuenca del ro Conchos, Mexico. Pro ceedings of the XX Congreso Nacional de Hidraulica, Toluca, Estado de Mexico. Asociacion Mexicana de Hidraulica, Mexico. ez, D., Mun oz, C. A., Reyes-Gomez, V. M., Nun Rodguez-Pineda, J. A., Gadsden, H., Lopez, M. and Hinojosa, O., 2006. Caracterisation de la secheresse ` hydrologique dans le bassin versant de la riviere Conchos, etat de Chihuahua, Mexique. Secheresse, 17(4). 475 484. Rodrguez-Pineda, J. A., Goodell, P., Reyes, I., Ren, M., oz, C., Hernandez, O. and Cruz, R., 2005. Mun Ciencia y tecnologa aplicada en el descubrimiento del origen de la contaminacion de agua subterranea con arsenico en Aldama, Chihuahua, Mexico. Memorias Taller Nacional Preparatorio al IV Foro Mundial del Agua, Cuernavaca, Mexico. 29 35. Schipper, E. L. F., 2007. Climate Change Adaptation and Development: Exploring the Linkages. Tyndall Centre Working Paper No. 107. Tyndall Centre for Climate Change Research, Norwich, UK. Trueba, V. and Goicochea, J., 2008. Escenarios del Agua 2015 2030 en la Cuenca del Ro Conchos: Acciones para un desarrollo ambientalmente sostenible. WWF-FGRA, Mexico. Vorosmarty, C. J., Green, P., Salisbury, J. and Lammers, R. B., 2000. Global water resources: vulnerability from climate change and population growth. Science, 289(5477). 284 288. Water Treaty, 1944. Tratado entre el gobierno de los Estados Unidos Mexicanos y el gobierno de Los Estados Unidos de America de la distribucion de las aguas inter nacionales de los ros Colorado, Tijuana y Bravo, desde Fort Quitman, Texas, hasta el Golfo de Mexico. portal. sre.gob.mx/cilanorte/pdf/1944.pdf. World Bank, 2007. El manejo del agua en territorios ind genas en Mexico [Water Management in Indigenous Lands in Mexico], 4(1). WWF, 2002. Estudio de Economa del Agua en los Distritos de Riego del Ro Conchos, Chihuahua. Consultor Arturo Puente G. Programa Desierto de Chihuahua. Convenio. QQ97-WWF-UK. WWF, 2004. Al Rescate del Ro Conchos [To the rescue of the Rio Conchos]. Internal report.

CLIMATE AND DEVELOPMENT

case study

Participatory river basin management in the Sao Joao River, Brazil: A basis for climate change adaptation?
LUIZ FIRMINO MARTINS PEREIRA1, *, SAMUEL BARRETO2 and JAMIE PITTOCK3
1

cio Ferreira, Av Getulio Vargas, 603 305/305, Centro, Araruama, 28970-000, Consorcio Intermunicipal Lagos Sao Joao, Edif

RJ, Brazil 2 lia, Brazil WWF Brazil, SHIS EQ QL 6/8, Conjunto E-2 andar, 71620-430, Bras
3

Fenner School of Environment & Society, Australian National University, Canberra ACT 0200, Australia

This paper describes an empirical case study of enhanced water management in the Sao Joao River basin on the southeast coast of Brazil between 1999 and 2008. The autonomous adaptation measures applied are assessed to derive lessons for more effective climate change adaptation. In response to severe eutrophication of their coastal lakes, effective, local multi stakeholder institutions were established under the auspices of the Consorcio Intermunicipal Lagos Sao Joao from 1999 to improve basin management. Having signicantly reduced the pollution problem, other environmental challenges are now being addressed. In managing environmental problems with multiple causes and effects, engaging multiple stakeholders and communicating the need to change environmental management, these local institutions have established the types of capacities needed for climate change adaptation. Factors contributing to the strengthening of this adaptive capacity include: engagement of local non-governmental organizations, companies and municipal governments; leadership and development of a collective identity; enabling national and state water laws; an ability to raise funds; and implementation of an iterative, adaptive management approach to environmental management. Keywords: adaptation; Brazil; climate change; institutions; river; Sao Joao; water

1. Introduction
This paper examines what motivated the commu nity in the Sao Joao region on the coast of southeastern Brazil (see Figure 1) to change their management of the basin and the factors that sustained these processes. It is published here as part of a special edition providing an overarching assessment (Pittock, 2009) of the global lessons derived from the six WWF empirical case studies of autonomous adaptation to climate change in developing countries. We start by reviewing relevant climate change impacts and adaptation concepts before detailing the situation in the Sao Joao region. The Intergovernmental Panel on Climate Change (IPCC) states that observational records

and climate projections provide abundant evidence that freshwater resources are vulnerable and have the potential to be strongly impacted by climate change (Bates et al., 2008). However, this paper looks at a region where there is very little published literature on climate change impacts or adaptation measures. Climate change forecasts for this area lack high resolution; however, impacts are expected from more extreme events (Pezza and Simmonds, 2005; Dufek and Ambrizzi, 2008), higher temperatures, sea level rise, possible increases in precipitation, exacerbation of water pollution, and loss of biodiversity (Christensen et al., 2007; Magrin et al., 2007; Bates et al., 2008). Even without climate change the world faces grave challenges in sustaining adequate water

B *Corresponding author. E-mail: Firmino@lagossaojoao.org.br


CLIMATE AND DEVELOPMENT 1 (2009) 261268 doi:10.3763/cdev.2009.0026 # 2009 Earthscan ISSN: 1756-5529 (print), 1756-5537 (online) www.earthscanjournals.com

262 Pereira, Barreto and Pittock

FIGURE 1 Location of the Sao Joao River basin (dark grey) and adjacent catchments (light grey) in the Lagos Sao Joao hydrographic region

resources, and the water sector has long applied adaptive management practices. The IPCC recognize this in saying (Kundzewicz et al., 2007, p. 196): Adaptation to changing conditions in water availability and demand has always been at the core of water management. The IPCC dene this type of autonomous adaptation (Bates et al., 2008, p. 48) as those that do not constitute a conscious response to climate stimuli, but result from changes to meet altered demands, objectives and expectations. We contend that there is much to be learnt for more effective adaptation from these measures. This paper also considers the benets of such

interventions in terms of increasing resilience and reducing vulnerability (Bates et al., 2008). We believe that the lack of high resolution climate change forecasts requires governments and societies to begin adaptation despite the uncertainties (Richardson et al., 2009). An institutional assessment by Tompkins et al. (2008) of disaster risk management and long-term adaptive capacity building identied four critical factors that led to reductions in risk: exible, learningbased, responsive governance (such as stakeholder participation, access to knowledge, accountability and transparency); committed, reform-minded and politically active actors; disaster risk reduction

CLIMATE AND DEVELOPMENT

Participatory river basin management in the Sao Joao River, Brazil 263

integrated into other social and economic policy processes; and a long-term commitment to managing risk. Tompkins and Adger (2004) argue that community-based management enhances adaptive capacity in two ways, by building networks that are important for coping with extreme events and by retaining the resilience of the underpinning resources and ecological systems. In the related eld of river basin governance, similar mechanisms have been proposed for enhancing water resources management, including communities of practice and social learning at different scales and involving diverse stakeholders (PahlWostl et al., 2007); as well as leadership and building collective identities (Abers, 2007). The IPCC propose a number of adaptation approaches to cope with uncertainty on climate change impacts. These include no-regrets policies that would generate net social and/or economic benets irrespective of whether or not anthropogenic climate change occurs, the increased use of water management measures that are relatively robust to uncertainty, and integrated water resources management (Bates et al., 2008). The potential impacts of and Brazils options for responding to climate change have been debated in the literature, but this work has largely focused on the Amazon and semi-arid northeast portions of the country, and on issues such as carbon balances, agriculture and biofuel production. Government policy offers little guidance. The National Plan on Climate Change (Government of Brazil, 2008) identies seven goals, of which two focus on adaptation. The proposed adaptation actions largely involve further research and communication activities, although strengthening of environmental sanitation measures is also proposed. In the Sao Joao River and adjacent coastal basins the biodiverse Atlantic forest remnants give way to farm lands, oodplains and coastal lagoons. The basin falls within the territories of 12 local governments in Rio de Janeiro State. In the 3,825 km2 region the resident population of 451,000 people swells to approximately 2 million people in holiday periods. The Juturnaiba Dam on the

120 km-long Sao Joao River is the main water supply for local people (Dantas et al., 2001). By the late 1990s expanding tourism development resulted in the coastal lagoons silting up and becoming polluted with untreated sewerage, causing a collapse in the shing industry and impacting on tourism. The Sao Joao basin was chosen for this autonomous adaptation case study because of: (a) WWFs long history of work in the region, starting in the 1960s to conserve an endangered primate, the Golden Lion Tamarin, which then led to the establishment of a freshwater conservation programme from 1999; and (b) the reforms in basin management since 1999.

2. Assessment
This paper reviews the changes in management of the Sao Joao national hydrographic region from 1999 to 2008 based on research undertaken in late 2008. The study sought to derive lessons concerning (unplanned) autonomous adaptation to climate change, with reference to the success factors for more effective adaptation and river basin management proposed by Tompkins and Adger (2004), Adger et al. (2005), Abers (2007) and Pahl-Wostl et al. (2007). We applied a largely qualitative analytical framework developed by Pittock (2009: Annex) to assess (a) autonomous adaptation, (b) socio-economic and (c) conservation outcomes. The assessment covers the local implementation of the 1997 national water law and 1999 Rio de Janeiro State water law (ANA, 2007) to decentralize and democratize water management (Brannstrom, 2004; Abers et al., 2006). In 1999 the Consorcio Intermunicipal Lagos Sao Joao (the Consortium) was formed by the 12 local governments and now includes four stakeholder representatives from the Sao Joao Basin Committee (Bidegain, 2002). This Committee was established in 2004 with membership from three tiers of government, academics, local companies and 58 civil society groups to engage basin residents more broadly and advise the Consortium (Pereira, 2007). Extensive investment in an

CLIMATE AND DEVELOPMENT

264 Pereira, Barreto and Pittock

environmental education programme from 2003 continues to build public support for catchment management reforms (Kobata, 2006). The Consortium established a process of developing and implementing a basin management plan (Bidegain and Pereira, 2006) with subsidiary work plans. The Consortium is now on its third work plan for the coastal lagoons. The Consortium secured resources from (a) secondment of a staff member from the state government to lead the secretariat; (b) membership fees from municipal governments scaled to reect the resident populations; and (c) participation fees from local companies. Establishment of a number of sub-basin and thematic working groups from 2005 has facilitated widespread participation in adaptive basin management, increasing local capacities. These institutions were established for integrated river basin management, to progressively solve major environmental problems, starting with water pollution and sheries management.

3. Results
A number of the key benets resulting from the strengthened institutions are summarized here in terms of adaptation, livelihood and environmental outcomes. The degradation of the rivers, Juturnaiba reservoir (30 km2), and Araruama (220 km2) and Saquarema (24 km2) coastal lagoons by discharge of untreated waste waters threatened the tourism and shing industries which comprise 70% of the regions economy. The Consortium, shing community and allied NGOs lobbied and took legal action against the state government pollution regulator. A key outcome was the renegotiation of water supply company concessions that saw an initial USD38.5 million investment in 2002 2005 in new sewerage treatment infrastructure that has reduced wastewater discharge by 75%. A USD19.3 million second phase is due to collect all waste waters for collection by 2009, and a third phase from 2010 to 2023 is planned to separate storm water from sewerage. In addition, the silted up entrance to the Araruama Lagoon

was dredged to restore greater exchange of water with the sea. The substantial reduction of pollution inows has reduced the threat that eutrophication of lagoon waters would be exacerbated by higher temperatures with climate change. Substantial socio-economic benets in restoring the shing and tourism industries have resulted from the interventions. Improved water quality has seen restoration of mangrove habitats and increases in sh, shrimp and bird populations. The shing industry has been re-established and now supports 600 families, and the tourism industry has recovered. Economic growth is increasing regional training and employment opportunities. The Consortium has also targeted disadvantaged sectors of the local society who often reside on and farm the most ood-prone lands. For instance, women in two communities are participating in a project to produce handicrafts for sale to tourists as a means of increasing and diversifying incomes to reduce poverty and their communities vulnerability to extreme events. Following success in reducing water pollution, the Committee and Consortium decided to scale up work from 2007 to reduce erosion and conserve the water sources and biodiversity through linking and restoring remnant riparian and other wetland habitats. The Juturnaiba Dam will be retrotted with a sh ladder at a cost of USD400,000 to reconnect populations of migratory species like grey mullet, sea bass and prawns, and the dams operating rules are being revised. The river bypass canal downstream of the dam will be decommissioned at a cost of USD700,000 to restore the Rio Sao Joaos natural course and the adjacent ood plain wetlands. The canal will be converted to aquaculture ponds, further diversifying the local economy. A payment for environmental services scheme is funding previously unemployed residents to restore riparian forests. This is reducing erosion and linking remnant habitats of a threatened primate, the Golden Lion Tamarin, whose population is increasing as the forests are restored. A network of protected areas is being established on private and public lands. Biodiversity and the

CLIMATE AND DEVELOPMENT

Participatory river basin management in the Sao Joao River, Brazil 265

shing industry are expected to benet further as reconnection and restoration of habitat increases species populations, access to habitat, ability to move to new habitats, and thus resilience to climate change impacts. As part of this assessment, local residents were asked their views on what made the institutional adaptation successful. Mr Arnaldo Villa Nova, President of the non-governmental organizations involved in the Consortium explained (pers. commun.): The Lagos Region has a wonderful natural heritage, which the civil society has always tried to preserve. [. . .We started the ght in] the year 2000, when we started with an indenite horizon, and with many doubts and questions to be answered. . . . How long will it resist degradation, deforestation, and economic exploitation without any scruples? When can our children play without the risk of getting sick? What can we do to recuperate and preserve the region? . . . The strategy applied by WWF in the Lagos Region was for direct support to the NGOs, which allowed the studies, investigations, the planning of projects for environmental recuperation, environmental education activities, as well as giving a denitive structure to the Consortium to accomplish its mission. After some years, the situation today is very different from the one in the beginning of the program. Many activities were implemented to stop degrading environmental processes, and increasing awareness of local government as well as the residents of the region. . . . We are half way in our journey: there is a lot to do . . .. Ms Denise Pena, a non-government representative on the Consortium focused on environmental education, commented (pers. commun.): The proposal of this new model of environmental management, where decisions must be taken by those who are acting and living in the territory of the hydrographic basin, in a decentralized and essentially participative form, could not take root without processes of

environmental education, which give conditions for productions and acquisition of knowledge, abilities and the development of attitudes, aiming at an individual and collective participation in this adopted model of management.

4. Discussion
In assessing the changes in the Sao Joao region we have identied a number of factors inuencing the success and the sustainability of the measures undertaken.

4.1. Motivation for change


The collapse of the coastal lagoon environments and consequent socio-economic impacts on the shing and tourism industries was the initial motivation for reform. The progress in Sao Joao appears to have been aided by funding provided by WWF to help local NGOs build their capacity, provide environmental education and develop a collective identity. This is consistent with Abers (2007) assessment of other Brazilian river basins. With respect to the debate over whether adaptation is better facilitated by focusing on social and biophysical risk reduction or by development to reduce poverty and enhance livelihoods (Adger, 2006; Schipper, 2007; Tompkins et al., 2008), in the case of Sao Joao, the primary investments enhanced livelihoods in the shing and tourism sectors as well as reducing physical vulnerability. Later and smaller-scale investments sought to improve the livelihoods of other disadvantaged groups, including through employment in environmental restoration. The Consortiums staff say that community awareness raising and engagement, and a virtuous and iterative cycle of successful interventions, has led to community support for further actions. This is consistent with the conclusion of Pahl-Wostl et al. (2007) that social learning institutions are vital, and Dovers (2005) who identied iterative programme cycles as being an element of

CLIMATE AND DEVELOPMENT

266 Pereira, Barreto and Pittock

successful sustainability policies. It is also consistent with Tompkins et al. (2008) who argue that stakeholder participation, access to knowledge, accountability and transparency are central to building long-term adaptive capacity.

4.2. Sustainability and funding


Institutional sustainability of these measures is enhanced by the local community engagement, mandate from the national and state water laws, and the fundraising capacity of the Consortium (Mea, 2007). While municipal and company fees do not pay all programme costs, they do enable leverage of other funds, including the secondment of state government staff. The basin insti tutions in Sao Joao differ from those in other parts of Brazil as they combine downward accountability through the leading role of municipal governments in the administrative Consortium, together with multi-stakeholder participation through the advisory council. This appears to compare favourably with three other institutional models for decentralized water resources management in Brazil (Brannstrom, 2004) in terms of promoting reform, limiting conicts, maximizing community engagement and accountability. This highlights the importance of concurrent measures across geopolitical scales, in this case at the individual, basin, local, state and federal government levels, for effective adaptation (Adger et al., 2005). The management interventions undertaken thus far appear to address some but not all likely impacts of climate change. The likelihood of algal blooms with warmer weather has diminished greatly with the extension of wastewater treatment, and further benets for aquatic biodiversity and sheries, reduced erosion and water quality are likely from the restoration of riparian forests, the construction of a sh ladder and removal of the channelized section of the river. Furthermore, some of the regions poorest communities have higher and more diverse incomes, enabling them to cope better with disruptive events. On the other hand, little

thought has yet been given to management of more frequent high rainfall events or to likely rises in sea level. However, the strength of the community-based management institutions supports Tompkins and Adgers (2004) proposition that greater adaptive capacity has been established through stronger social networks and by retaining the resilience of the underpinning resources and ecological systems. The problems dealt with by the river basin management institutions to date have the same attributes as those of climate change adaptation challenges: multiple cause and effect linkages, multiple stakeholders and communication of the need for changes in environmental management. Having addressed eutrophication of regional water bodies and now riparian restoration, there is the capacity and will in the basin institutions to manage the new problems expected to come with climate change.

4.3. Barriers and lessons


Until this study commenced, the basin management institutions had not considered how to manage climate change. They appeared discouraged by the uncertainties in data available on the likely local impacts of climate change and lack of locally available expertise. The Consortium staff saw the climate change information available to them as lacking salience (Meinke et al., 2006). As a result of this research, the Consortium staff are now inspired to reassess how their programme can now become more climate informed, including by implementing further no-regrets adaptation measures. This case study highlights the importance of strong local institutions for adaptation. The extensive public communication and engagement has made government institutions more accountable and responsive (Costa, 2007). The multi-stakeholder Committee and Consortium processes built partnerships and consensus for change, and stopped buck-passing between governments. This is consistent with the systematic social learning promoted by Lee (2003). The

CLIMATE AND DEVELOPMENT

Participatory river basin management in the Sao Joao River, Brazil 267

Consortium secretariat was kept small and work was contracted out to other institutions in the basin, enhancing engagement, partnerships and capacities for reform. The basin institutions subsidiarity mechanisms enhanced local ownership of problems, innovation and successful responses, consistent with Abers (2007).

interventions by achieving substantial early successes. B Proponents of mainstreaming climate change adaptation must communicate in salient language and illustrate no- and low-regrets options that are effective despite uncertainties as to climate change impacts. The reforms at Sao Joao highlight the opportunities to mainstream climate change adaptation through river basin management programmes.

4.4. Potential to scale up


Brazils national and state water laws could enable similar work in the approximately 140 similar river basin institutions across Brazil. Abers (2007) and Brannstrom (2004, p. 231) outline factors that have favoured or hindered decentralized stakeholder governance in other Brazilian basins, and by comparison the Sao Joao institutional framework appears to encourage a threeway dynamic among central authorities, local government and civil society and develop a common local identity particularly effectively.

Acknowledgements
HSBC contributed funding to WWF Brazils par ticipation in the Sao Joao project, as well as to this assessment.

References
Abers, R. N., 2007. Organizing for governance: building collaboration in Brazilian river basins. World Development, 35(8). 1450 1463. Abers, R. N., Formiga-Johnson, R. M., Frank, B., Keck, M. E. and Lemos, M. C., 2006. Stakeholder Councils and River Basin Management in Brazil: Democratizing Water Policy? III Encontro da ANPPAS, 23 26 May 2006. Braslia-DF. Adger, W. N., 2006. Vulnerability. Global Environmental Change, 16(3). 268 281. Adger, W. N., Arnell, N. W. and Tompkins, E., 2005. Successful adaptation to climate change across scales. Global Environmental Change Part A, 15(2). 77 86. ANA, 2007. Water Resources Management in Brazil. http:// hidroweb.ana.gov.br/cd2/water/docs/part2.htm. Bates, B. C., Kundzewicz, Z. W., Wu, S. and Palutikof, J. P. (eds), 2008. Climate Change and Water. Technical paper of the Intergovernmental Panel on Climate Change. IPCC Secretariat, Geneva. Bidegain, P., 2002. Lagoa de Araruama Perl do Maior Ecossistema Lagunar Hipersalino do Mundo. Semads, Rio de Janeiro. Bidegain, P. and Pereira, L. F. (eds), 2006. Plano de Bacia Hidrograca da Regiao dos Lagos e do Rio Sao Joao, Tomo I, II e III. Rio de Janeiro Comite de Bacia Lagos Sao Joao Consorcio Intermunicipal Lagos Sao Joao. Brannstrom, C. (2004). Decentralizing water resource management in Brazil. European Journal of Development Research, 16(1). 214 234.

5. Conclusions
Management of the Sao Joao basin did not consider climate change, but the institutional reforms and other interventions have established a strong basis for building resilience and reducing vulnerability. A number of lessons can be drawn from this case for more effective adaptation to climate change:
B

Severe pollution of the regions water bodies helped mobilize non-governmental organizations and local leaders to respond. Three factors were crucial to the success of national and state river-basin management institutions in facilitating reform at the basin scale: bringing together diverse stakeholders to work towards a common vision; local ownership; and an independent nancing mechanism. Concurrent investment in activities that both reduced vulnerability and enhanced livelihoods generated community support and inspired community condence in new

CLIMATE AND DEVELOPMENT

268 Pereira, Barreto and Pittock

Christensen, J. H., Hewitson, B., Busuioc, A., Chen, A., Gao, X., Held, I., Jones, R., Kolli, R. K., Kwon, a W.-T., Laprise, R., Magan Rueda, V., Mearns, L., Menendez, C. G., Raisanen, J., Rinke, A., Sarr, A. and Whetton, P., 2007. Regional Climate Projections. Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, S. Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor and H. L. Miller (eds). Cambridge University Press, Cambridge, UK and New York. 847 940. Costa, L. B., 2007. Social Transformation: When Participation and Empowerment are Not Enough. Participatory Management in Lagos Sao Joao River Basin, RJ, Brazil. MSc Report in Development Studies Specialization: Politics of Alternative Development. Institute of Social Studies, The Netherlands. Dantas, M. E., Shinzato, E., Medina, A. I. M., Silva, C. R., Pimentel, J., Lumbreras, J. F., Calderano, S. B. and Carvalho Filho, A., 2001. Diagnostico Geoambiental do Estado do Rio de Janeiro, Braslia/CPRM. Dovers, S., 2005. Environment and Sustainability Policy: Creation, Implementation, Evaluation. Federation Press, Annandale, VA. Dufek, A. S. and Ambrizzi, T., 2008. Precipitation varia bility in Sao Paulo State, Brazil. Theoretical and Applied Climatology, 93(3 4). 167 178. Government of Brazil, 2008. Executive Summary. National Plan on Climate Change (English version). Interministerial Committee on Climate Change. Government of Brazil, Brasilia. Kobata, C., 2006. A Educacao Ambiental como ferramenta da gestao participativa na Microbacia do Rio Cambucaes, Silva Jardim, RJ. Dissertacao apresentada ao Curso de Pos-Graduacao em Ciencia Ambiental da Universidade Federal Fluminense, como requisito parcial para obtencao do Grau de Mestre. Area de Concentracao: Analise de Processos Socio-Ambien tais. Universidade Federal Fluminense, Niteroi. Kundzewicz, Z. W., Mata, L. J., Arnell, N. W., Doll, P., Kabat, P., Jimenez, B., Miller, K. A., Oki, T., Sen, Z. and Shiklomanov, I. A., 2007. Freshwater resources and their management. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, M. L. Parry, O. F. Canziani, J. P. Palutikof, P. J. van der Linden and C. E. Hanson (eds). Cambridge University Press, Cambridge, UK. 173 210. Lee, K. N., 1993. Compass and Gyroscope: Integrating Science and Politics for the Environment. Island Press, Washington, DC and Covela, CA.

Magrin, G., Gay Garca, C., Cruz Choque, D., Gimenez, J. C., Moreno, A. R., Nagy, G. J., Nobre, C. and Villamizar, A., 2007. Latin America. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. M. L. Parry, O. F. Canziani, J. P. Palutikof, and P. J. van der Linden and C. E. Hanson (eds). Cambridge University Press, Cambridge, UK. 581 615. Mea, R. D., 2007. Payment for Environmental Services in the Atlantic Forest: A Feasible Solution to Deforestation? MSc Report in Environmental Policy. Imperial College London, Faculty of Natural Sciences. Meinke, H., Nelson, R., Kokic, P., Stone, R., Selvaraju, R. and Baethgen, W., 2006. Actionable, climate knowledge: From analysis to synthesis. Climate Research, 33. 101 110. Pahl-Wostl, C., Craps, M., Dewulf, A., Mostert, E., Tabara, D. and Taillieu, T., 2007. Social learning and water resources management. Ecology and Society, 12(2). 5. Pereira, L. F. M., 2007. A gestao participativa no caso do saneamento da Regiao dos Lagos, Rio de Janeiro. Revista Discente Expressoes Geogracas. Florianopolis-SC, 03. 10 41. Pezza, A. B. and Simmonds, I., 2005. The rst South Atlantic hurricane: unprecedented blocking, low shear and climate change. Geophysical Research Letters, 32(15). 5. Pittock, J., 2009. Lessons for climate change adaptation from better management of rivers. Climate and Development, 1(3). 194 212. doi: 10.3763/cdev.2009. 0021. Richardson, K., Steffen, W., Schellnhuber, H. J., Alcamo, J., Barker, T., Kammen, D. M., Leemans, R., Liverman, D., Munasinghe, M., Osman-Elasha, B., Stern, N. and Waever, O., 2009. Synthesis Report. Climate Change: Global Risks, Challenges and Decisions. University of Copenhagen, Copenhagen. Schipper, E. L. F., 2007. Climate Change Adaptation and Development: Exploring the Linkages. Tyndall Centre Working Paper No. 107. Tyndall Centre for Climate Change Research, Norwich, UK. Tompkins, E. L. and Adger, W. N., 2004. Does adaptive management of natural resources enhance resilience to climate change? Ecology and Society, 9(2). 10. Tompkins, E. L., Lemos, M. C. and Boyd, E., 2008. A less disastrous disaster: managing response to climatedriven hazards in the Cayman Islands and NE Brazil. Global Environmental Change Human and Policy Dimensions, 18(4). 736 745.

CLIMATE AND DEVELOPMENT

review article

Embracing uncertainty in freshwater climate change adaptation: A natural history approach


JOHN H. MATTHEWS* and A. J. WICKEL
Conservation Science Department, World Wildlife Fund, 1250 24th Street, NW, Washington, DC 20037, USA

Climate shifts are not new in the experience of humans and other species, but the capacity of potential evolutionary and ecological responses to climate change has been reduced through widespread human modications of natural ecosystems. The magnitude, duration and timescales of altered climate threats require multigenerational strategies for climate change adaptation. In many places terrestrial and aquatic species and human livelihoods are limited by the availability of freshwater resources. Current climate change adaptation practice places great faith in the ability of climate models to predict specic impacts, which then become the focus of climate change adaptation activities and thus foster reactive impacts thinking. Given that freshwater climate variables are associated with high predictive uncertainty, a novel approach referred to here as adaptation thinking treats ecosystems as dynamic entities that will be inherently different from current and past ecosystem states for multiple reasons, including climate change. As a result, adaptation thinking emphasizes the shifting relationship between institutions and ecosystems. This approach promotes exibility and continuous scenario development. Using natural modes of adaptation as a template for sustainable development should promote collaboration between scientists, policymakers and development professionals. Keywords: climate change; climate change adaptation; conservation; economic development; freshwater

1. Introduction
The Fourth Assessment Report of the Intergovernmental Panel on Climate Change (IPCC) projects that even with immediate implementation of climate mitigation policies, the global climate system will continue to shift and change for decades. Two centuries of industrial emissions of greenhouse gases have altered the radiative forcing of the planet, resulting in rapid changes in what humans perceive as stationary climate conditions (IPCC, 2007a; Biggs et al., 2009; CCSP, 2009). Thus, the need to adjust and adapt to realized shifts in climate and to prepare for major shifts in what we have perceived as normal weather is crucial for individuals, institutions, ecosystems and species. Climate change adaptation (CCA) will be with us for the foreseeable future. While it may be of some comfort that humans and other species have adapted to changes in climate conditions many times in the past, the explicit recognition of the importance of treating climate as a variable rather than as a constant factor is new for humans. Evolutionary ecologists believe that past dramatic large-scale climate change events have had three levels of impacts on the life history of organisms: alterations in the geographic range and phenological1 patterns of species, species extinctions, and (over long timescales) shifts in the processes of natural selection and speciation (Parmesan and Yohe, 2003). The current period of climate change may qualify as such a serious large-scale event (Root et al., 2003, Marris, 2007), particularly given the synergy of climate change to massive shifts in land use that have also occurred over the past

B *Corresponding author. E-mail: john.matthews@wwfus.org


CLIMATE AND DEVELOPMENT 1 (2009) 269279 doi:10.3763/cdev.2009.0018 # 2009 Earthscan ISSN: 1756-5529 (print), 1756-5537 (online) www.earthscanjournals.com

270 Matthews and Wickel

150 years (e.g. Dahl, 1990; Ricciardi and Rasmussen, 1999; Abell, 2002).

2. Climate change adaptation as a natural process


Despite a growing consensus that humans need to adapt to climate change, there is little agreement about the actual meaning of CCA (Brooke, 2008; Williams et al., 2008). The IPCC dened CCA as an adjustment in ecological, social, or economic systems in response to actual or expected climatic stimuli and their effects or impacts . . . to changes in processes, practices, or structures to moderate or offset potential damages or to take advantage of opportunities associated with changes in climate (IPCC, 2001). This denition does not distinguish between the spatial or temporal scale of climate impacts nor the severity of those impacts, even though such considerations could result in fundamentally different approaches to CCA that should be explicitly addressed by conservation and economic development groups. Fundamental to the concept of CCA is the implication of vulnerability to negative impacts from climate change, which must be assessed through some process before responding with appropriate adaptations. Vulnerability is described by the IPCC as having three components: sensitivity to change, exposure to climate shifts, and degree of adaptive capacity (IPCC, 2001, Williams et al., 2008). Signicantly, Brooke (2008) points out that most CCA practitioners focus on either human adaptation (economic development) or species and ecosystem adaptation (conservation) with much less emphasis on a more holistic sustainability science that incorporates both development and conservation, such as through linked social-ecological systems (Holling, 1973). The means of presenting climate data through institutions such as the IPCC may be indirectly inuencing how policymakers and development and conservation groups perceive or frame shifts in climate. For instance, much current CCA

work is rmly grounded on deterministic assessments of climate change impacts, primarily derived from climate models. Such CCA approaches thus effectively assume that (a) climate shifts will be gradual, (b) shifts in mean annual climate variables are more important than the frequency or severity of extreme weather events, (c) high-condence CCA plans can be dened on the basis of circulation model projections for regional climate decades from now, (d) that the temporal scale of circulation model resolution is appropriate for describing ecosystem-level impacts, and (e) that ecosystems will remain largely intact and recognizable as assemblages of species in the face of climate, or will shift as ecosystem units. Many of these perspectives reect data presentation or model bias rather than climate or ecological processes. Making these assumptions would lead to CCA projects that are most akin to the biological denition of acclimatization, i.e. adjustments by an individual organism to changed conditions (Futuyma, 1998); often these adjustments have a specic end point or target in mind. For instance, to cope with increased frequency of droughts, a resource manager might lower water demand levels to a more sustainable level through increased water-use efciency. The livelihood, species or ecosystem is stretched but not fundamentally recongured (in the history of the discipline of biology, such a response might be considered a Lamarckian model of evolution). Climate change here is well bounded as a problem, whose impacts can be known and solved. In many cases, the (often) unstated assumptions about climate science are made against the explicit advice of climate modellers (e.g. IPCC, 2008). Moreover, most of these assumptions are not well founded on ecological science, especially for freshwater ecosystems (Allen and Ingram, 2002; Poff et al., 2002; Parmesan, 2006, Hall et al., 2008). Perhaps the most important underlying concept behind CCA projects that rely on most or all of these assumptions is that uncertainty surrounding future climate and emerging ecosystem impacts can be reasonably constrained

CLIMATE AND DEVELOPMENT

A natural history approach 271

or reduced (Bammer and Smithson, 2008). Climate change science does not create highcondence quantitative projections at local spatial scales for most climate variables. Worse, ecology (the basis for most conservation work) is much more of a historical, descriptive science like geology than a non-historical, predictive science such as physics (Hilborn and Mangel, 1997; Pilkey and Pilkey-Jarvis, 2008). Even if we knew what the future climate was going to look like, we are unlikely accurately to capture the panoply of ecosystem and species impacts. The majority of climate models typically describe climate change as gradual shifts in mean climate (IPCC, 2001). Such a mode of climate change differs from instances in the paleo-climate record that show changes in the frequency of extreme weather events (the number of tropical storms, intense precipitation events, drought severity/duration). Likewise, past climates frequently changed in abrupt shifts from one climate plateau to another plateau (Anderson et al., 2007; CCSP, 2009). Gradual shifts are only one means of climate transition. Change is an old theme in evolutionary biology, however. In Darwinian evolutionary ecological usage adaptation refers to multigenerational changes, which implies developing genetic tness with a species environment over time (i.e. survival of the ttest). An adaptation may begin incrementally but typically evolves to become very different from some initial condition (Futuyma, 1998). Adaptation with a natural history perspective emphasizing the environmental setting of an organism probably does not have a designated end point; change is a process, and the direction of change itself may shift. A resource manager may experiment in series with drought-response systems, such as a completely new irrigation delivery network, changes in crop selection, and new planting and harvesting methods, while also developing new systems to monitor drought severity and frequency patterns and creating new stakeholder institutions to negotiate for reduced ows. Such an approach emphasizes adaptation as a process extending across decades

or generations, involving experiments and tradeoffs, learning and (certainly) failures. Here, climate change presents obstacles, but these targets are both moving and associated with uncertainty. Viewing adaptation from an evolutionary ecological perspective situates CCA into the broader engagement between humans and other species with climate over long timescales (Diamond, 2004; Fagan, 2008). Focusing on CCA as a natural process also leads to an examination of the aspects of climate that are most likely to be inuencing the evolutionary ecology of non-human species and, by extension, of human societies, cultures and livelihoods. The literature on impacts and responses to climate change has often focused on air temperature, a trend reinforced by the widespread usage of the term global warming. Although temperature is an important determinant of physical habitat, the availability and predictability of freshwater resources is likely to be a far more relevant aspect of climate to species and humans, natural ecosystems and agriculture (Poff et al., 2002; Parmesan and Galbraith, 2004). Indeed, we believe that if carbon is the key target for climate mitigation work, then freshwater should be the key focus for CCA, for both terrestrial and aquatic biomes. Given this assumption, CCA for aquatic species and humans and most other terrestrial species will be dominated by the story of impacts on water quality, water quantity, and the timing of ows and hydroperiod (Poff et al., 2002; Nilsson and Renofalt, 2008; Palmer et al., 2008; Poff, 2009). We wish to advocate here an approach to CCA that is process based, that realistically accounts for the high uncertainties associated with the emerging climate, and that grounds both species conservation and economic development in the reality of shifting ecosystem conditions and qualities. Fundamentally, we are suggesting that humans be viewed as a species like other species, which implies that we exist within ecological boundaries and on an evolutionary trajectory. An evolutionary ecological perspective is not particularly controversial in the context of conservation biology, but we believe the implications of this approach have not been effectively

CLIMATE AND DEVELOPMENT

272 Matthews and Wickel

communicated to policymakers or institutions working in the eld of economic development. As practising scientists, we believe that these perspectives have important policy implications for freshwater resources and the livelihoods dependent on them, given their importance to economies, ecosystem services and livelihoods. The gap between conservation and economic development must be bridged rapidly for CCA to be successful globally and tailored locally. Applying these assumptions and perspectives to the natural history (i.e. the environmental setting and context) of human societies has the potential to improve the long-term viability and relevance of economic development. These issues are not academic or trivial. If CCA represents a new path for relating human institutions to ecosystems, then we risk wasting limited conservation and development resources unless we adopt the correct approach to our work. Perhaps most signicantly, we risk reducing the inherent capacity to adapt to climate shifts (i.e. maladaptation) through climatically unsustainable resource management (Pittock and Dovers, 2009; Pittock, 2009). Here, we focus on a series of issues about CCA for freshwater resources from a scientic perspective relevant to emerging climate change issues in policy, resource management and sustainable development practice: what shifts should occur in conservation and economic development practices in a CCA context? How do we manage freshwater resources for a dynamically shifting climate? What is an appropriate model for successful CCA projects?

3. Moving beyond stationarity towards adaptive freshwater management


Until very recently, almost all water resource management practice has assumed that the best basis for infrastructure design and management was captured through the historical record of that basins hydrological variability an assumption of ecosystem stationarity. Many basins, particularly in western Europe and North America, have

long historical records (.100 years) of discharge based on monitoring stations. But the basic assumption that recent knowledge served as an effective guide to the future was not widely questioned. More recently, stationarity has been declared dead as a result of human-induced climate change (Milly et al., 2008). While climate models are reasonably good at describing and predicting air temperature trends, many other climate variables do not have similar levels of condence (Milly et al., 2005; Nohara et al., 2006). The analysis of historical trends and projections from circulation models suggest that some regional climate regimes are shifting into new states that are substantially different from what have long been viewed as normal climate states. The Murray Darling basin in Australia, for instance, is either in a historically unprecedented drought or in a new and unfamiliar climate regime, depending on the authors perspective (Pittock, 2003). The rapid loss of tropical glaciers in the Himalayas (Kehrwald et al., 2008) likewise suggests climate regimes that are signicantly different for humans and other species in these regions. In extreme cases, emerging hydrological regimes may represent so-called no-analog climates that is, they will be profoundly different from what has been seen over the past several millennia (Fox, 2007). Thus, the recent past will serve as an increasingly less reliable guide to the future. In theory, projections of future climate could be detailed enough to provide high-condence predictions of what regional or local climate will look like in a particular place at a particular time. These projections could then guide infrastructure development. A growing body of technical literature has been developed to describe the process of downscaling circulation and hydrological models from large spatial scales to guide particular projects and planning. Unfortunately, this route is fraught with risk for planners and policymakers (Fowler and Wilby, 2007). The dominant components that govern the water balance precipitation and evapotranspiration are extremely variable in both space and time. These components do not show robust,

CLIMATE AND DEVELOPMENT

A natural history approach 273

high-condence trends within individual circulation models and show even less agreement when comparing different circulation models. Indeed, even if the models themselves were very accurate, we do not know how future economic conditions, technological shifts, or mitigation policies (scenarios, in IPCC parlance) will alter the concentrations of greenhouse gases in the atmosphere. These levels of uncertainty increase substantially as spatial scale decreases (from regional to local) or temporal period grows more distant (farther into the future) (IPCC, 2008). The lack of certainty represents a serious crisis for water resource management and planning. The death of stationarity means the future is (probably) cloudier and less certain (Milly et al., 2008). Downscaling models may be inappropriate and generate false condence. At worst, they may result in development and conservation practices that are maladaptive. Developing appropriate guidelines for their use is challenging (Fowler and Wilby, 2007; Johnson and Weaver, 2009). The implications for this crisis in hydrology and water engineering are profound for those involved in freshwater-related economic development and conservation. Water resource management institutions and individuals working in both areas have generally been consumers of hydrological data, and thus we too have implicitly assumed that stationarity is alive and well, whether or not we were aware that we held that assumption. In response, we can (a) proceed as if nothing has changed, (b) allow the lack of clarity and condence in what we know about freshwater ecosystems to excuse inaction, or (c) begin to modify our work in a way that takes account of climatic and eco-hydrological uncertainty. This third path represents the new element of freshwater climate adaptation relative to traditional conservation and development.

sustainable resource management: conservationists have attempted to reduce pressures on ecosystems, while development advocates focused on improving resource allocation equity and efciency. In many parts of the world, both groups are in competition for scarce nancial resources and the attention of policymakers and electorates. In the worst cases, conservation may even be focused on excluding resources from human communities that already exist on a precarious economic cliff. Indeed, human-caused climate change has the potential to worsen this relationship. But these divisions have always been articial and distracting, and climate change simply increases the urgency with which we must reduce conict. Instead, we wish to emphasize that climate change also presents an opportunity to discuss sustainable resource management in a more holistic way, bringing together opposing groups and focusing policymakers drawn to concerns over climate change on a coherent view of sustainable development (e.g. Ebert et al., 2009). The priorities from conservation and development groups have too often differed even when they shared the same goals, and climate change offers a chance to unify our work and make both perspectives more effective over coming decades. Ultimately, such a union represents a means of addressing future-climate uncertainty, which is the common enemy of all. A natural history approach suggests a process-oriented focus that begins with natural systems and actively depends on human monitoring, exible management and a dynamic relationship between economies and ecosystems. In a perfect world, the ideal goal of CCA would be to anticipate what impacts will happen in particular localities at particular times to particular species or livelihoods. This goal is threatened by the difculty of anticipating impacts, across at least three levels of uncertainty:

4. Understanding and constraining climate uncertainty


Conservation biologists and development professionals have differed in their approaches to 1. variables of climate in model projections estimating future climate, which are used to estimate . . .

CLIMATE AND DEVELOPMENT

274 Matthews and Wickel

2. . . . inows and outows for freshwater ecosystems in model projections estimating future impacts on freshwater ecosystems, which are used to estimate . . . 3. . . . synergistic impacts between such factors as climate change, development trajectory and freshwater ecosystems. Levels 2 and 3 form the primary basis for developing most CCA plans (e.g. Williams et al., 2008), but they are also the levels most burdened with high levels of uncertainty. These models can be developed generally using two approaches: an analysis of climate trends at a particular locality, or downscaled circulation models and associated eco-hydrological models. Trend analyses are normally a robust means of predicting short-term impacts at local scales, but for statistical power they require several decades of reliable data, which is often elusive. The applicability of trend analyses over multi-decadal timescales is probably low. The levels of uncertainty for circulation models are even higher for watery variables, such as the seasonal timing of precipitation, the amount of precipitation for a particular region, or whether that precipitation will fall in liquid or frozen forms. Relative to air temperature variables, precipitation and evapotranspiration variables are extremely difcult to determine in long-term circulation models (Milly et al., 2005; Nohara et al., 2006, Koutsoyiannis et al., 2008). Both approaches make numerous assumptions about how local or regional climate functions. It is critical for policymakers and non-science development staff to recognize that while scientic models are approximations of real systems, climatic and eco-hydrological models cannot be downscaled to the functionally local scale with the condence necessary for most conservation and development projects. Thus, while they may be informative and a guide to action, they are unlikely to be denitive. The problem is less of technical or computational challenges than of high levels of stochasticity and variability in the behaviour of precipitation in weather and climate systems. Finally, it is worth reiterating

that project-based work must avoid assuming a new stationarity, as if climate change is directed towards some new stable plateau. Human-made climate change is driving major shifts in freshwater ecosystems, giving conservation and development practitioners a moving target on a scale ranging from decades to centuries (IPCC, 2007a,b). A seemingly stable assessment of water quantity in a particular lake in 2020, for instance, may not be representative of that lake a few decades later. And a moving target will be more difcult to hit. Placing an unduly high degree of condence in such models tends to lead to an emphasis on responding to some specic set of predicted impacts on ecosystems or livelihoods what could be called impacts thinking. This approach is widespread and dominates much of the conservation and economic development discussion about CCA at this time. Impacts thinking suggests that CCA strategies can be clearly articulated, that models are robust and comprehensive, and that adaptation does not represent a signicant shift in the worldview behind resource management. Impacts thinking often implies that vulnerability need only be assessed once.

5. From impacts thinking to adaptation thinking


Adaptation thinking shifts the focus of CCA from a particular set of ecosystem or livelihood impacts to the process of resource management itself. The challenges to successful CCA for resource management are developing a productive means of responding to impacts uncertainty over time. This task has three major components: 1. understanding what qualities enable local ecosystems (and in many cases traditional livelihoods) to autonomously adapt and remain resilient to climate impacts; 2. understanding how resource management institutions can facilitate (or at least not inhibit) these processes;

CLIMATE AND DEVELOPMENT

A natural history approach 275

3. developing the institutional means to anticipate and detect processes of climate-driven change, as well as to implement responses to realized and potential impacts. These three components of constraining uncertainty are relatively well known and understood. Much contemporary hydro-ecological theory, for instance, focuses on how lakes, rivers and wetlands retain ecosystem health (and, for humans, reliable ecosystem services) such as mitigating the impacts of surrounding land-use shifts, overabstraction of water, restoring geomorphological structure, and promoting ecological connectivity within and across ecosystems (Abell, 2002). This body of knowledge stocks the toolbox for CCA. The tool users are the focus of components 2 and 3 the resource managers, development institutions and policymakers rather than scientists per se. These two components are also relatively well understood paths: they require that CCA proceed as a process rather than a single event and that major decisions (such as oodplain restoration, building dams, irrigation infrastructure or selecting climate-appropriate crops) are based on a risk assessment basis on the best-available scientic knowledge. Thus, uncertainty, vulnerability and risk are evaluated continuously or iteratively. Component 3 becomes the means for developing robust monitoring and climate trend and ecosystem impact detection systems that can re-evaluate the degree of uncertainty and risk. In other words, high levels of uncertainty in decision-making processes suggest that exibility is the appropriate response until that uncertainty has been reduced. If water resources appear to be declining relative to demand, constructing a new dam may be appropriate, but reducing demand through increased efciency may be the more exible route until more certainty exists about the degree of water availability shift. After all, in most countries it would be difcult to tear down a dam once it has been constructed, much less undo the potential negative ecosystem consequences of dam construction and management. In contrast, a focus on demand should be at least in theory

more elastic and less expensive through techniques such as increasing efciency, depending on a wider variety of regional freshwater sources, and developing drought-management plans. Component 3 will also require a strong relationship between scientic, development and policy realms so that analytical approaches can balance model condence with policy and governance priorities. Taken together, adaptation thinking represents a novel approach to the existing conservation and economic development work practised globally. By shifting the focus on CCA from ecosystem and livelihoods impacts to the process of determining appropriate resource management, adaptation thinking requires policy to ultimately focus on building effective and adaptive governance structures and institutions. Water utilities that can actively monitor and manage throughout a district become a powerful means of reducing the impacts of droughts on both people and ecosystems. Agriculture departments that advocate a basket of less water-intensive crops in regions with increased precipitation variability can mitigate the increasing demand of a rapidly growing population. There can be no doubt that more dams will be necessary in the future. But dams may not be the only solution, and they often do not assist in solving more than a small number of narrowly dened problems. An overdependence on hard concrete solutions such as dams and other capital-intensive infrastructure to what are really institutional problems may ultimately be undermined by the original problem of stationarity: can the new infrastructure perform well enough with a shifting climate in order to realistically provide a return on the original investment? In other words, expensive infrastructure will often be a part of the suite of decisions contemplated for water-related problems, but the dilemma of climate uncertainty also means that the ows may be insufcient to generate expected levels of electricity, increased evapotranspiration may remove large quantities of stored water, or dam heights may be incapable of dealing with increasingly extreme precipitation events. Maintaining exibility is a

CLIMATE AND DEVELOPMENT

276 Matthews and Wickel

powerful approach to a wide range of climate-informed decisions over long timescales.

6. Conclusions: Turning climate uncertainty into an advantage


Flexibility in and of itself is a useful overarching theme for CCA. However, exibility as applied to a more specic set of principles is more directive and operationally satisfying when planning new projects. Based on the experience of the World Wildlife Fund (WWF) in recent years, we advocate the following set of concepts for creating climate-informed sustainable development work:

6.1. Focus on institutional capacity for implementing and responding to emerging climate conditions
Too often, infrastructure management has been a source of maladaptation rather than a tool for improving CCA processes (Pittock, 2009). Regional and local water resource management institutions should, in most cases, be the unit of focus for both economic development and freshwater conservation work (Burton, 1996; see also Ebert et al., 2009; Gujja et al., 2009; Pereira et al., 2009). Fundamentally, they must be capable of developing climate-appropriate policies and of implementing those policies, such as working effectively with farmers, irrigation districts, shers and other socio-economic groups (see Roux et al., 2008). But these institutions must also move from a model of organizational stationarity to become climate-adaptive institutions that are capable of detecting changes in relevant ecosystems, evaluating the appropriateness of their relationship with those ecosystems, and shifting their behaviour as more effective models emerge.

Knight, 1998; Parmesan et al., 1999; Easterling et al., 2000; IPCC, 2008). As a result, many regions are experiencing both more oods and more droughts, while tropical storm activity is probably being inuenced by climate change as well. Since climate change may be creating more severe versions of these than have been experienced for decades or centuries, existing emergency management plans may be inadequate. If current planning focuses on so-called 100-year oods, for instance, this standard may be statistically based on the past century but inaccurate when applied to oods for the next 10 or 20 years, which will dene a new standard. Drought length and severity may prove especially challenging in regions such as the southeastern US when they face water shortages requiring locally unusual responses such as rationing. Some of the pain of previously inexperienced extremes can be reduced by sound disaster management planning with policymakers. When disasters do occur, they represent opportunities to engage the interest of stakeholders and policymakers in climate change issues even though they also present the risk of creating policy that is reactive rather than thoughtful (Barrios et al., 2009). However, they also present the risk of reinforcing inequitable allocations and derailing development.

6.3. Consider impacts at regional and basin levels, even when water-management decisions are made locally
Unfortunately, eco-hydrological surface, atmospheric or groundwater networks do not normally align with political boundaries or institutional zones of inuence. Thus, local freshwater issues are rarely local in a hydrological sense (e.g. see Barrios et al., 2009; Kashaigili et al., 2009). Without consideration of regional and basin impacts from local decisions, a problem that one community has solved may in fact have simply shifted to another community downstream.

6.2. Negotiate responses to extreme weather and climate variability before crises occur
Globally, climate change is creating more variability in extreme weather events (Karl and

CLIMATE AND DEVELOPMENT

A natural history approach 277

6.4. Develop effective ecosystem and livelihood monitoring and analysis systems
One of the key assumptions of adaptation thinking is that the projections of impacts on ecohydrological systems are fraught with high levels of uncertainty. As a result, accurate and timely monitoring and detection of shifts in key variables in water quality, quantity and timing are not perfect means to capture emerging trends (especially major state-level shifts), but they can track gradual shifts and, ideally, help anticipate tipping points. Such systems need not be expensive or centrally organized, but they do need to be standardized, effective and robust (Pittock, 2009; Yu et al., 2009). They would need to focus concurrently on climate impacts and human impacts on water resources with a clear understanding of their differences in scale local, regional, national or international. They should also be afliated with institutions that are capable of analysing trends and explaining those trends to the stakeholders who need to use this knowledge.

Acknowledgements
We are grateful to Jamie Pittock for sponsoring our paper in this issue, and to our colleagues globally at WWF who have been an inspiration and motivation to explore the practical problems of climate adaptation in a freshwater context.

Note
1. Phenologies that are time- or season-sensitive behaviours, such as the onset of blooming or the initiation or cessation of long-distance bird migration.

References
Abell, R., 2002. Conservation biology for the biodiversity crisis: a freshwater follow-up. Conservation Biology, 16(5). 1435 1437. Allen, M. R. and Ingram, W. J., 2002. Constraints on future changes in climate and the hydrologic cycle. Nature, 419. 224 232.

Anderson, D., Goudie, A. and Parker, A., 2007. Global Environments Through the Quaternary. Oxford University Press, Oxford, UK. Bammer, G. and Smithson, M., 2008. Uncertainty and Risk: Multidisciplinary Perspectives. Earthscan Press, London. Barrios, J. E., Rodrguez-Pineda, J. A. and De la Maza, M., 2009. Integrated river basin management in the Conchos River basin, Mexico: a case study of freshwater climate change adaptation. Climate and Development, 1(3). 249260. doi:10.3763/cdev.2009.0024. Biggs, R., Carpenter, S. and Brock, W., 2009. Turning back from the brink: detecting an impending regime shift in order to avert it. Proceedings of the National Academy of Sciences, 106(3). 826 831. doi:10.1073/pnas.0811729106. Brooke, C., 2008. Conservation and adaptation to climate change. Conservation Biology, 22(6). 14711476. Burton, I., 1996. The growth of adaptation capacity: practice and policy. Adapting to Climate Change: An International Perspective, J. Smith, N. Bhatti, G. Menzhulin, R. Benioff, M. I. Budyko, M. Campos, B. Jallow and F. Rijsberman (eds). Springer-Verlag, New York. 55 67. CCSP (Climate Change Science Program), 2009. Thresholds of Climate Change in Ecosystems. Synthesis and Assessment Product 4.2. D. Fagre and C. Charles (lead authors). US Geological Survey, Reston, VA. Dahl, T. E., 1990. Wetland Losses in the United States: 1780s to 1980s. US Department of the Interior, Fish and Wildlife Service, Washington, DC. Diamond, J., 2004. Collapse: How Societies Choose to Fail or Succeed. Viking Press, New York. Easterling, D. R., Evans, J. L., Groisman, P. Ya., Karl, T. R., Kunkel, K. E. and Ambenje, P., 2000. Observed variability and trends in extreme climate events: a brief review. Bulletin of the American Meteorological Society, 81. 417 425. Ebert, S., Hulea, O. and Strobel, D., 2009. Floodplain restoration along the lower Danube: a climate change adaptation case study. Climate and Development, 1(3). 212 219. doi:10.3763/cdev.2009.0022. Fagan, B., 2008. The Great Warming: Climate Change and the Rise and Fall of Civilizations. Bloomsbury Press, London. Fowler, H. and Wilby, R., 2007. Editorial: beyond the downscaling comparison story. International Journal of Climatology, 27. 1543 1545. Fox, D., 2007. Back to the no-analog future? Science, 316. 823 825. Futuyma, D., 1998. Evolutionary Biology (3rd edn). Sinaeur Associates, Sunderland, MA. Gujja, B., Dalai, S., Shaik, H. and Goud, V., 2009. Adapting to climate change in the Godavari River basin of

CLIMATE AND DEVELOPMENT

278 Matthews and Wickel

India by restoring traditional water storage systems. Climate and Development, 1(3). 229 240. doi:10.3763/cdev.2009.0020. Hall, N., Stuntz, B. and Abrams, R., 2008. Climate change and freshwater resources. Natural Resources and Environment, 22(3). 30 35. Hilborn, R. and Mangel, M., 1997. The Ecological Detective. Princeton University Press, Princeton, NJ. Holling, C. S., 1973. Resilience and stability of ecological systems. Annual Review of Ecology and Systematics, 4. 1 23. IPCC (Intergovernmental Panel on Climate Change), 2001. Climate Change 2001: Impacts, Adaptation, and Vulnerability, Contribution of Working Group II to the Intergovernmental Panel on Climate Change Third Assessment Report, J. J. McCarthy, O. F. Canziani, N. A. Leary, D. J. Dokken and K. S. White (eds). Cambridge University Press, Cambridge, UK. IPCC (Intergovernmental Panel on Climate Change), 2007a. Fourth Assessment Report. Climate Change 2007: The Physical Science Basis, S. Solomon, D. Qin, M. Manning, M. Marquis, K. Averyt, M. M. B. Tignor, H. L. Miller and Z. Chen (eds). Cambridge University Press, Cambridge, UK. IPCC (Intergovernmental Panel on Climate Change), 2007b. Fourth Assessment Report. Climate Change 2007: Impacts, Adaptation, and Vulnerability, M. Parry, O. Canziani, J. Palutikoff, P. van der Linden and C. Hanson (eds). Cambridge University Press, Cambridge, UK. IPCC (Intergovernmental Panel on Climate Change), 2008. Climate Change and Water, B. Bates, Z. Kundzewicz, S. Wu and J. Palutikoff (eds). Technical Paper of the Intergovernmental Panel on Climate Change, IPCC Secretariat, Geneva. Johnson, T. and Weaver, C., 2009. A framework for assessing climate change impacts on water and watershed systems. Environmental Management, 43. 118 134. Karl, T. R. and Knight, R. W., 1998. Secular trends of precipitation amount, frequency, and intensity in the United States. Bulletin of the American Meteorological Society, 79(2). 231 242. Kashaigili, J. J., Rajabu, K. and Masolwa, P., 2009. Freshwater management and climate change adaptation: experiences from the Great Ruaha River catchment in Tanzania. Climate and Development, 1(3). 220228. doi:10.3763/cdev.2009.0025. Kehrwald, N., Thompson, L., Tandong, Y., MosleyThompson, E., Schotterer, U., Almov, V., Beer, J., Eikenberg, J. and Davis, M., 2008. Mass loss on Himalayan glacier endangers water resources. Geophysical

Research Letters, 35. L22503. doi:10.1029/2008GL0 35556. Koutsoyiannis, D., Efstradiadis, A., Mamassis, N. and Christodes, A., 2008. On the credibility of climate predictions. Hydrological Sciences, 53(4). 671 684. Marris, E., 2007. What to let go. Nature, 450. 152 155. Milly, P. C. D., Dunne, K. A. and Vecchia, A. V., 2005. Global pattern of trends in streamow and water availability in a changing climate. Nature, 438. 347 350. Milly, P., Betancourt, J., Falkenmark, M., Hirsch, R., Kundzewicz, Z., Lettenmaier, D. and Stouffer, R., 2008. Stationarity is dead: whither water management? Science, 319. 573 574. Nilsson, C. and Renofalt, B., 2008. Linking ow regime and water quality in rivers: a challenge to adaptive catchment management. Ecology & Society, 13(2). 18. Nohara, D., Kitoh, A., Hosaka, M. and Oki, T., 2006. Impact of climate change on river runoff. Journal of Hydrometeorology, 7. 1076 1089. Palmer, M., Reidy Liermann, C., Nilsson, C., Florcke, M., Alcamo, J., Lake, P. and Bond, N., 2008. Climate change and the worlds river basins: anticipating management options. Frontiers in Ecology and the Environment, 6(2). 81 89. Parmesan, C., 2006. Evolutionary and ecological responses to recent climate change. Annual Reviews of Ecology, Evolution, and Systematics, 37. 637 669. Parmesan, C. and Galbraith, H., 2004. Observed Impacts of Global Climate Change in the U.S. Pew Center on Global Climate Change, Arlington, VA. Parmesan, C. and Yohe, G., 2003. A globally coherent ngerprint of climate change impacts across natural systems. Nature, 421. 37 42. Parmesan, C., Root, T. L. and Willig, M., 1999. Impacts of extreme weather climate on terrestrial biota. Bulletin of the American Meteorological Society, 81(3). 441 450. Pereira, L. F. M., Barreto, S. and Pittock, J., 2009. Partici patory river basin management in the Sao Joao River, Brazil: a basis for climate change adaptation? Climate and Development, 1. doi:10.3763/cdev.2009.026. Pilkey, O. and Pilkey-Jarvis, L., 2008. Useless Arithmetic: Why Environmental Scientists Cant Predict the Future. Columbia University Press, New York. Pittock, B., 2003. Climate Change: An Australian Guide to the Science and Potential Impacts. Australian Greenhouse Ofce, Canberra. Pittock, J., 2009. Lessons from climate change adaptation for better management of rivers. Climate

CLIMATE AND DEVELOPMENT

A natural history approach 279

and Development, 1(3). 194 211. doi:10.3763/cdev. 2009.0021. Pittock, J. and Dovers, S., 2009. Why a special issue on adaptation and water management? Climate and Development, 1(3). 191193. doi:10.3763/cdev.2009. 0019. Poff, N. L., 2009. Managing for variability to sustain freshwater ecosystems. Journal of Water Resources Planning and Management, Jan/Feb. 1 4. Poff, N. L., Brinson, M. M. and Day, J. W., 2002. Aquatic Ecosystems and Global Climate Change: Potential impacts on Inland Freshwater and Coastal Wetland Ecosystems in the United States. Pew Center on Global Climate Change, Arlington, VA. Ricciardi, A. and Rasmussen, J. B., 1999. Extinction rates of North American freshwater fauna. Conservation Biology 13(5). 1220 1222.

Root, T., Price, J., Hall, K., Schneider, S., Rosenzweig, C. and Pounds, J., 2003. Fingerprints of global warming on wild animals and plants. Nature, 421. 57 60. Roux, D., Ashton, P., Nel, J. and MacKay, H., 2008. Improving cross-sector policy integration cooperation in support of freshwater conservation. Conservation Biology, 22(6). 1382 1387. Williams, S., Shoo, L., Isaac, J., Hoffmann, A. and Langham, G., 2008. Towards an integrated framework for assessing vulnerability of species to climate change. PLoS Biology, 6(12). e325. Yu, X., Jiang, L., Wang, J., Li, L., Wang, L. and Lei, G., 2009. Freshwater management and climate change adaptation: experiences from the central Yangtze in China. Climate and Development, 1(3). 241 248. doi:10.3763/cdev.2009.0023.

CLIMATE AND DEVELOPMENT

Das könnte Ihnen auch gefallen