20.04.2014 Views

Coastal vulnerability, resilience and adaptation to ... - OceanDocs

Coastal vulnerability, resilience and adaptation to ... - OceanDocs

Coastal vulnerability, resilience and adaptation to ... - OceanDocs

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

COASTAL VULNERABILITY, RESILIENCE AND<br />

ADAPTATION TO CLIMATE CHANGE<br />

AN INTERDISCIPLINARY PERSPECTIVE<br />

Richard J.T. Klein<br />

Kumulative Dissertation


This thesis has been submitted <strong>to</strong> the Mathematisch-Naturwissenschaftliche Fakultät of the<br />

Christian-Albrechts-Universität zu Kiel in part-fulfilment of the requirements for the degree<br />

of Dr. rer. nat.<br />

December 2002<br />

Referent: Prof. Dr. Horst Sterr<br />

Ko-Referent: Prof. Dr. Ir. Pier Vellinga<br />

Cover pho<strong>to</strong>: Cley-next-the-Sea, United Kingdom.<br />

Pho<strong>to</strong>graph taken by Richard Klein in June 1996.


This PhD thesis is dedicated <strong>to</strong> Michiel, Ma <strong>and</strong> Ruth.<br />

Thank you for your inspiration <strong>and</strong> love, for being there as long as it lasted. I wish I could<br />

share the joy of completion with you. You will always be in my heart.


Preface <strong>and</strong> acknowledgements<br />

This thesis is the result of about eight years of sometimes more, sometimes less focused research<br />

on a subject that <strong>to</strong>ok a long time <strong>to</strong> be defined as the <strong>to</strong>pic of a PhD. This thesis does<br />

not originate from a single, well-defined PhD project but rather combines <strong>and</strong> integrates results<br />

of a number of different studies that at first seemed only loosely related at best. However,<br />

some three years ago a common theme emerged, which also happened <strong>to</strong> be most<br />

timely in the scientific <strong>and</strong> policy debates on climate change. It is increasingly recognised<br />

that the concept of <strong>vulnerability</strong> <strong>to</strong> climate change encompasses more than the exposure <strong>and</strong><br />

sensitivity of natural <strong>and</strong> human systems <strong>to</strong> potential impacts of climate change; it is also defined<br />

by the degree <strong>to</strong> which these systems can prepare for <strong>and</strong> respond <strong>to</strong> impacts.<br />

The common theme in my work of the past eight years has been the analysis of elements<br />

that, one way or another, determine or describe this latter part of <strong>vulnerability</strong> <strong>to</strong> climate<br />

change. My research has aimed <strong>to</strong> contribute <strong>to</strong> answering the question as <strong>to</strong> how coastal systems<br />

<strong>and</strong> communities would <strong>and</strong> could respond <strong>to</strong> climate change <strong>and</strong>, in particular, how<br />

this response may be assessed as part of coastal <strong>vulnerability</strong> studies. The focus on coastal<br />

zones was not a conscious decision but it has turned out <strong>to</strong> be a fortunate one, as coastal<br />

zones provide an ideal “labora<strong>to</strong>ry” for developing <strong>and</strong> testing new conceptual ideas. Compared<br />

<strong>to</strong> other effects of climate change, sea-level rise is relatively straightforward in that<br />

the direction of change is known. Moreover, it is a fairly certain consequence of climate<br />

change, which has well-documented analogues in recent geological his<strong>to</strong>ry.<br />

For the same reason that this thesis did not start off from a clearly defined question or<br />

hypothesis it does not have well-defined results or conclusions. This thesis is <strong>to</strong> be considered<br />

very much as “work in progress” <strong>and</strong> indeed, a range of initiatives are currently being developed<br />

that use concepts <strong>and</strong> information presented in this thesis <strong>and</strong> elsewhere as a starting<br />

point or framework for their analysis. I find it very exciting <strong>to</strong> see the way in which my work<br />

contributes <strong>to</strong> the development of what have been referred <strong>to</strong> as “second-generation <strong>vulnerability</strong><br />

studies”, also in sec<strong>to</strong>rs other than coastal zones. Second-generation <strong>vulnerability</strong><br />

studies promise <strong>to</strong> provide a much stronger link between science <strong>and</strong> policy than the previous<br />

generation of studies could.<br />

The research presented in this thesis has been conducted at the following three institutes:<br />

the Institute for Environmental Studies of the Vrije Universiteit Amsterdam, the School of Environmental<br />

Sciences of the University of East Anglia <strong>and</strong> the Potsdam Institute for Climate<br />

Impact Research. In addition, I have benefited from the many interactions with the Flood<br />

Hazard Research Centre of Middlesex University. I am proud not only that I have worked with<br />

some of the brightest minds in the field but also that some close friendships have developed.<br />

I hope there will be many more opportunities <strong>to</strong> work <strong>to</strong>gether <strong>and</strong> cultivate our friendships.<br />

I am very grateful <strong>to</strong> everybody who, over the course of the past eight years, has stimulated<br />

me <strong>to</strong> do the work that I have done <strong>and</strong> convinced me not <strong>to</strong> be discouraged by the initial<br />

lack of focus, the difficulties in securing research funding or the fact that my work does<br />

not have an empirical basis. I would like <strong>to</strong> mention the following people in particular.<br />

Pier Vellinga introduced me <strong>to</strong> the work on climate change <strong>and</strong> coastal zones, particularly<br />

in relation <strong>to</strong> the Intergovernmental Panel on Climate Change. Pier motivated me <strong>to</strong> develop<br />

<strong>and</strong> explore my initial ideas <strong>and</strong> gave me the courage <strong>to</strong> share them with a network of<br />

distinguished yet approachable international scientists. Robert Nicholls in particular has been<br />

a wonderful sounding board for my ideas, always willing <strong>to</strong> discuss them, preferably over a<br />

drink. These discussions often turned out <strong>to</strong> be a fruitful basis for joint publications <strong>and</strong> have<br />

thus been instrumental in shaping my thoughts <strong>and</strong> developing the focus I needed <strong>to</strong> complete<br />

my PhD. Ian Bateman <strong>and</strong> Kerry Turner have added an environmental economics dimension <strong>to</strong><br />

my work <strong>and</strong> have generated my strong interest in interdisciplinary research. I am grateful for<br />

the time they were willing <strong>to</strong> invest in me <strong>and</strong> I cherish the memories of my stay in Norwich.<br />

iv


Perhaps without knowing it, Richard Tol, by being his provocative self, has always stimulated<br />

<strong>and</strong> in a way empowered me <strong>to</strong> pursue my own ideas. Whilst at times we disagree about<br />

the use of different analytical methods <strong>and</strong> the interpretation of results, I greatly value our<br />

mutual respect <strong>and</strong> willingness <strong>to</strong> learn from each other. I would like <strong>to</strong> thank John Schellnhuber<br />

<strong>and</strong> Carlo Jaeger for being so supportive of my work <strong>and</strong> for encouraging me <strong>to</strong> finalise<br />

this thesis. They have created a highly stimulating working environment in which I have the<br />

opportunity <strong>to</strong> continue <strong>and</strong> exp<strong>and</strong> my research on environmental <strong>vulnerability</strong> <strong>and</strong> <strong>adaptation</strong>.<br />

Finally, I could not have wished for a better PhD supervisor than Horst Sterr, whose constructive<br />

criticism is certainly not the reason why it <strong>to</strong>ok me longer than anticipated <strong>to</strong> complete<br />

my PhD. I thank him for his motivating style of supervision <strong>and</strong>, above all, his patience.<br />

I am also grateful <strong>to</strong> all those with whom I have had the privilege <strong>and</strong> pleasure <strong>to</strong> work<br />

over the past eight years <strong>and</strong> who have been willing <strong>to</strong> share their ideas so as <strong>to</strong> help me <strong>to</strong><br />

develop mine. These include Neil Adger, Joe Alcamo, James As<strong>to</strong>n, Pieter van Beukering,<br />

Luitzen Bijlsma, Luke Br<strong>and</strong>er, Nick Brooks, Earle Buckley, Ian Bur<strong>to</strong>n, Michele Capobianco,<br />

Alex<strong>and</strong>er Carius, Brian Challenger, Dave Dokken, Peter Dok<strong>to</strong>r, Kees Dorl<strong>and</strong>, Tom Downing,<br />

Kris Ebi, Bud Ehler, Frank Eierdanz, Jan Feenstra, Hasse Goosen, Poul Grashoff, Dolf de<br />

Groot, Joyeeta Gupta, John H<strong>and</strong>mer, Madeleen Helmer, Frank Hoozemans, Cees Hulsbergen,<br />

Saleemul Huq, Venugopalan Ittekkot, Roger Jones, Arun Kashyap, Mick Kelly, Philipp Knill,<br />

Sari Kovats, Alcira Kreimer, Dörthe Krömker, Onno Kuik, Kavi Kumar, Liza Leclerc, Neil Leary,<br />

Rik Leemans, Bo Lim, Holger Lip<strong>to</strong>w, Don MacIver, Marcel March<strong>and</strong>, Ajay Mathur, Elisabeth<br />

Mausolf, Roger McLean, Bettina Menne, Bert Metz, Ben Mieremet, Nobuo Mimura, Robbert<br />

Misdorp, Richard Moss, Isabelle Niang-Diop, Leonard Nurse, Brett Orl<strong>and</strong>o, Gualbert Oude<br />

Essink, Jean Palutikof, An<strong>and</strong> Patwardhan, Stephen Peake, Edmund Penning-Rowsell, Martha<br />

Perdomo, John Pernetta, Olga Pilifosova, Sachooda Ragoonaden, Agustin Sánchez-Arcilla, Ravi<br />

Sharma, Barry Smit, Joel Smith, Marcel Stive, Roger Street, Dennis Tänzler, Ferenc Tóth, Nassos<br />

Vafeidis, Bert van der Valk, Laura Van Wie-McGrory, Roda Verheyen, Jan Vermaat, Claudio<br />

Volonté, Richard Warrick, Robert Watson, Gary Yohe, as well as all my colleagues in Potsdam.<br />

Of my colleagues in Potsdam I would particularly like <strong>to</strong> thank Johann Grüneweg for translating<br />

the summary of this thesis in<strong>to</strong> German <strong>and</strong> Lilibeth Acosta-Michlik for scanning the published<br />

papers that form the core of this thesis.<br />

Financial support for my work has been kindly provided by the Dutch <strong>and</strong> German governments,<br />

the British Council, the European Union, the United Nations Development Programme,<br />

the United Nations Environment Programme, the Secretariat of the United Nations Framework<br />

Convention on Climate Change, the Disaster Management Facility of the World Bank, Delft<br />

Hydraulics, Norfolk Wildlife Trust, the Institute for Environmental Studies at the Vrije Universiteit<br />

Amsterdam <strong>and</strong> the Potsdam Institute for Climate Impact Research.<br />

Finally, I would like <strong>to</strong> thank my friends for putting up with me, for supporting me <strong>and</strong><br />

for being there when it mattered. You know who you are. I also thank my father <strong>and</strong> sister for<br />

being the best father <strong>and</strong> sister I have. And I thank Kate for being at the right place at the<br />

right time <strong>and</strong> for being there ever since. Poppy, I love you <strong>to</strong> bits!<br />

This PhD thesis is dedicated <strong>to</strong> three people who I miss every day <strong>and</strong> without whom I<br />

would not have been the person I am <strong>to</strong>day. Death <strong>and</strong> divorce hurt, yet they can never erase<br />

the wonderful memories I have. Michiel, Ma, Ruth, this is for you.<br />

Potsdam, December 2002<br />

Richard Klein<br />

v


I do not know what I may appear <strong>to</strong> the world; but <strong>to</strong> myself I seem <strong>to</strong> have been only like a<br />

boy playing on the seashore, <strong>and</strong> diverting myself now <strong>and</strong> then finding a smoother pebble or<br />

a prettier shell than ordinary, whilst the great ocean of truth lay all undiscovered before me.<br />

Sir Isaac New<strong>to</strong>n<br />

vii


viii


Table of contents<br />

Preface <strong>and</strong> acknowledgements<br />

Table of contents<br />

Klein, R.J.T., 2002: <strong>Coastal</strong> <strong>vulnerability</strong>, <strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong> <strong>to</strong> climate change:<br />

an introduc<strong>to</strong>ry overview.<br />

Klein, R.J.T. <strong>and</strong> R.J. Nicholls, 1999: Assessment of coastal <strong>vulnerability</strong> <strong>to</strong> climate<br />

change. Ambio, 28(2), 182–187.<br />

Klein, R.J.T., M.J. Smit, H. Goosen <strong>and</strong> C.H. Hulsbergen, 1998: Resilience <strong>and</strong><br />

<strong>vulnerability</strong>: coastal dynamics or Dutch dikes? The Geographical Journal, 164(3),<br />

259–268.<br />

Klein, R.J.T. <strong>and</strong> I.J. Bateman, 1998: The recreational value of Cley Marshes Nature<br />

Reserve: an argument against managed retreat? Water <strong>and</strong> Environmental<br />

Management, 12(4), 280–285.<br />

Smit, B., I. Bur<strong>to</strong>n, R.J.T. Klein <strong>and</strong> J. W<strong>and</strong>el, 2000: An ana<strong>to</strong>my of <strong>adaptation</strong> <strong>to</strong><br />

climate change <strong>and</strong> variability. Climatic Change, 45(1), 223–251.<br />

Klein, R.J.T., R.J. Nicholls <strong>and</strong> N. Mimura, 1999: <strong>Coastal</strong> <strong>adaptation</strong> <strong>to</strong> climate change:<br />

can the IPCC Technical Guidelines be applied? Mitigation <strong>and</strong> Adaptation Strategies<br />

for Global Change, 4(3–4), 239–252.<br />

Klein, R.J.T., R.J. Nicholls, S. Ragoonaden, M. Capobianco, J. As<strong>to</strong>n <strong>and</strong> E.N. Buckley,<br />

2001: Technological options for <strong>adaptation</strong> <strong>to</strong> climate change in coastal zones.<br />

Journal of <strong>Coastal</strong> Research, 17(3), 531–543.<br />

iv<br />

ix<br />

1<br />

32<br />

39<br />

50<br />

57<br />

87<br />

103<br />

Klein, R.J.T., 2002: Synthesis <strong>and</strong> next steps. 117<br />

Summary 127<br />

Deutsche Zusammenfassung 129<br />

Biography 133<br />

ix


<strong>Coastal</strong> Vulnerability, Resilience <strong>and</strong> Adaptation <strong>to</strong> Climate Change:<br />

An Introduc<strong>to</strong>ry Overview<br />

Richard J.T. Klein<br />

1. Introduction<br />

The work presented in this PhD thesis has not been carried out within a single, well-defined<br />

project. Instead, it integrates the results of a number of studies conducted from 1994 onwards,<br />

each of which had different clients <strong>and</strong> objectives. The common theme of the studies<br />

has been the description <strong>and</strong> analysis of elements that determine how coastal systems <strong>and</strong><br />

communities would <strong>and</strong> could respond <strong>to</strong> climate change <strong>and</strong>, in particular, how this response<br />

may be assessed as part of coastal <strong>vulnerability</strong> studies.<br />

<strong>Coastal</strong> zones are amongst the most dynamic natural environments on earth, providing a<br />

range of goods <strong>and</strong> services that are essential <strong>to</strong> human social <strong>and</strong> economic well-being.<br />

<strong>Coastal</strong> zones represent the narrow transitional zone between the world’s l<strong>and</strong> <strong>and</strong> oceans,<br />

characterised by highly diverse ecosystems such as coral reefs, mangroves, beaches, dunes<br />

<strong>and</strong> wetl<strong>and</strong>s. Many people have settled in coastal zones <strong>to</strong> take advantage of the range of<br />

opportunities for food production, transportation, recreation <strong>and</strong> other human activities provided<br />

here. A large part of the global human population now lives in coastal areas: estimates<br />

range from 20.6 per cent within 30 km of the sea <strong>to</strong> 37 per cent in the nearest 100 km <strong>to</strong> the<br />

coast (Cohen et al., 1997; Gommes et al., 1998; Nicholls <strong>and</strong> Small, 2002). In addition, a considerable<br />

portion of global economic wealth is generated in coastal zones (Turner et al.,<br />

1996). Many coastal locations exhibit a growth in population <strong>and</strong> income higher than their national<br />

averages (Carter, 1988; WCC’93, 1994), as well as substantial urbanisation (Nicholls,<br />

1995a).<br />

In the Second Assessment Report of the Intergovernmental Panel on Climate Change<br />

(IPCC), Bijlsma et al. (1996) noted that climate-related change in coastal zones represents<br />

potential additional stress on systems that are already under intense <strong>and</strong> growing pressure.<br />

The IPCC concluded that although the potential impacts of climate change by themselves may<br />

not always pose the greatest threat <strong>to</strong> natural coastal systems, in conjunction with other<br />

stresses they could become a serious issue for coastal societies, particularly in those places<br />

where the <strong>resilience</strong> of the coast has been reduced. This conclusion has been the main motivation<br />

behind my research, which has aimed at better underst<strong>and</strong>ing the <strong>vulnerability</strong>, <strong>resilience</strong><br />

<strong>and</strong> <strong>adaptation</strong> of coastal zones in the face of climate change.<br />

The insights gained in the various studies have been the basis of a number of peer-reviewed<br />

<strong>and</strong> published papers, six of which form the core of this PhD thesis. Each of these papers<br />

explores different aspects of coastal <strong>vulnerability</strong>, <strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong> <strong>to</strong> climate<br />

change. This introduc<strong>to</strong>ry chapter provides the context for coastal <strong>vulnerability</strong> assessment<br />

by giving an overview of current stresses in coastal zones, as well as of the possible effects of<br />

climate change on coastal sustainability. It also explores how the three concepts that form<br />

the basis of this PhD thesis (<strong>vulnerability</strong>, <strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong>) have been defined <strong>and</strong><br />

applied in other disciplines. Finally, this chapter defines the research objectives pursued in<br />

this thesis, outlines the methodological approach taken <strong>and</strong> introduces the six papers.<br />

Following the six peer-reviewed <strong>and</strong> published papers, a synthesis chapter aims <strong>to</strong> draw<br />

conclusions in the light of existing <strong>and</strong> emerging scientific <strong>and</strong> policy needs. The synthesis<br />

chapter also provides an agenda for research that can build on the findings of this PhD thesis.<br />

2. Sustainable development <strong>and</strong> global change in coastal zones<br />

This PhD thesis focuses on coastal <strong>vulnerability</strong>, <strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong> <strong>to</strong> climate change.<br />

However, climate change is not the only challenge <strong>to</strong> sustainable development in coastal<br />

zones. Other developments taking place on a global scale also have a pervasive but immediate<br />

impact on coastal sustainability. These developments are associated with the many eco-<br />

1


nomic opportunities offered by natural coastal systems. They include overexploitation of<br />

coastal resources, pollution, increasing nutrient fluxes, decreasing freshwater availability,<br />

sediment starvation <strong>and</strong> urbanisation. These non-climatic changes affect the natural capacity<br />

of coastal systems <strong>to</strong> cope with stresses such as climate change <strong>and</strong> therefore need <strong>to</strong> be considered<br />

along with the potential impacts of climate change on coastal zones.<br />

All natural coastal systems are—<strong>to</strong> varying degrees—able <strong>to</strong> offset the effects of human<br />

intervention. However, many coastal areas have been so heavily modified <strong>and</strong> intensively developed<br />

that their natural <strong>resilience</strong> <strong>to</strong> further changes has been substantially reduced (see<br />

Section 4.2). In many places, unsustainable development of coastal resources has progressively<br />

increased <strong>vulnerability</strong> <strong>to</strong> both the natural dynamics associated with present-day climate<br />

variability <strong>and</strong> anticipated impacts of global climate change.<br />

2.1. Functions of natural coastal systems<br />

Natural coastal systems (comprising both the geomorphic <strong>and</strong> ecological components) support<br />

a variety of socio-economic activities, including <strong>to</strong>urism <strong>and</strong> recreation, exploitation of living<br />

<strong>and</strong> non-living resources (e.g., fisheries, aquaculture, agriculture <strong>and</strong> extraction of water, oil<br />

<strong>and</strong> gas), industry <strong>and</strong> commerce, infrastructure development (e.g., ports, harbours, bridges,<br />

roads <strong>and</strong> coastal defence works) <strong>and</strong> nature conservation. In addition, natural coastal systems<br />

are crucial in safeguarding environmental quality. For example, they are important in<br />

assimilating waste, maintaining migration <strong>and</strong> nursery habitats—<strong>and</strong> thereby biodiversity—<strong>and</strong><br />

providing protection against extreme events. Thus, human well-being depends directly or indirectly<br />

on the availability of environmental goods <strong>and</strong> services provided by natural coastal<br />

systems.<br />

In general terms, environmental functions (defined by De Groot (1992a) as the capacity<br />

of the natural environment <strong>to</strong> provide goods <strong>and</strong> services that satisfy human needs in a sustainable<br />

manner) can be categorised as regulation functions, user <strong>and</strong> production functions<br />

<strong>and</strong> information functions (De Groot, 1992a; Vellinga et al., 1994). Table 1 gives an overview<br />

of all environmental functions identified in the literature. Most of these functions are relevant<br />

<strong>to</strong> coastal zones. The relative importance of these functions for a given coastal area depends<br />

on the ecological characteristics, socio-economic circumstances <strong>and</strong> management objectives<br />

of the area in question.<br />

Regulation functions<br />

Protection against harmful cosmic influences;<br />

Regulation of the local <strong>and</strong> global energy balance;<br />

Regulation of the chemical composition of the atmosphere;<br />

Regulation of the chemical composition of the oceans;<br />

Regulation of the local <strong>and</strong> global climate;<br />

Regulation of runoff, flood prevention;<br />

Water catchment, groundwater recharge;<br />

Prevention of soil erosion, sediment control;<br />

Formation of <strong>to</strong>psoil, maintenance of soil fertility;<br />

Fixation of solar energy, biomass production;<br />

S<strong>to</strong>rage <strong>and</strong> recycling of organic matter;<br />

S<strong>to</strong>rage <strong>and</strong> recycling of nutrients;<br />

S<strong>to</strong>rage <strong>and</strong> recycling of human waste;<br />

Regulation of biological control mechanisms;<br />

Maintenance of migration <strong>and</strong> nursery habitats;<br />

Maintenance of biological (<strong>and</strong> genetic) diversity.<br />

Information functions<br />

User <strong>and</strong> production functions<br />

Providing space <strong>and</strong> a suitable substrate for:<br />

– human habitation <strong>and</strong> (indigenous) settlements;<br />

– cultivation (crop growing, animal husb<strong>and</strong>ry, aquaculture);<br />

– energy conversion;<br />

– transportation <strong>and</strong> navigation;<br />

– recreation <strong>and</strong> <strong>to</strong>urism;<br />

– nature protection;<br />

Production of:<br />

– oxygen;<br />

– water (for drinking, irrigation, industry, etc.);<br />

– food <strong>and</strong> nutritious drinks;<br />

– genetic resources;<br />

– medicinal resources;<br />

– raw materials for clothing <strong>and</strong> household fabrics;<br />

– raw materials for building, construction <strong>and</strong> industrial use;<br />

– biochemicals (other than fuels <strong>and</strong> medicines);<br />

– fuel <strong>and</strong> energy;<br />

– fodder <strong>and</strong> fertiliser;<br />

– ornamental resources.<br />

Aesthetic information;<br />

Spiritual <strong>and</strong> religious information;<br />

His<strong>to</strong>ric information;<br />

Cultural <strong>and</strong> artistic inspiration;<br />

Scientific <strong>and</strong> educational information.<br />

Table 1 — Functions of the biosphere (adapted from De Groot, 1992a).<br />

Regulation functions are crucial in safeguarding environmental quality. They include the<br />

regulation of erosion <strong>and</strong> sedimentation patterns, which helps <strong>to</strong> prevent floods, the regulation<br />

of the chemical composition of the atmosphere <strong>and</strong> the oceans, which helps <strong>to</strong> control<br />

2


pollution, <strong>and</strong> the maintenance of migration <strong>and</strong> nursery habitats, which sustains biological<br />

diversity. Thus, many natural <strong>and</strong> semi-natural ecosystems play a fundamental part in the<br />

regulation of essential processes that contribute <strong>to</strong> the maintenance of a healthy environment<br />

<strong>and</strong> the long-term stability of the biosphere. In coastal zones, regulation functions are<br />

particularly important in maintaining <strong>resilience</strong> in the face of natural hazards such as s<strong>to</strong>rm<br />

surges (Nicholls <strong>and</strong> Branson, 1998) <strong>and</strong> epidemiological threats such as those posed by harmful<br />

algae blooms (Anderson <strong>and</strong> Garrison, 1997).<br />

User <strong>and</strong> production functions are essential in providing the many living <strong>and</strong> non-living<br />

resources that are utilised by society. These functions include the provision of space <strong>and</strong> a<br />

suitable substrate for human habitation, agriculture, recreation <strong>and</strong> <strong>to</strong>urism <strong>and</strong> the production<br />

of, among other things, food, raw materials, fuel <strong>and</strong> energy. <strong>Coastal</strong> resources are often<br />

of great commercial importance, yet their full exploitation can cause multiple-use conflicts<br />

amongst diverse stakeholders (Goldberg, 1994; Vallega, 1996).<br />

Information functions relate <strong>to</strong> the part that nature plays in meeting human intellectual<br />

<strong>and</strong> emotional needs. The scientific, his<strong>to</strong>rical, spiritual <strong>and</strong> aesthetic information derived<br />

from coastal ecosystems provides unique opportunities for cognitive development <strong>and</strong> emotional<br />

enrichment. In addition, the awareness of human society being part of a larger system<br />

forms an important basis for the notion that nature, including coastal ecosystems, also has a<br />

value of its own, independent of human needs or perceptions (Turner, 1988).<br />

Sustainable development in coastal zones is only attained when it enables the coastal<br />

system <strong>to</strong> self-organise, that is, <strong>to</strong> perform all its potential functions without adversely affecting<br />

other natural or human systems (cf. Klein et al., 1998). However, coastal zones have<br />

increasingly been subjected <strong>to</strong> the unsustainable use <strong>and</strong> unrestricted development of l<strong>and</strong><br />

<strong>and</strong> resources, with the sole aim of maximising the financial benefits provided by user <strong>and</strong><br />

production functions. This has resulted in an increasingly large area of cultivated <strong>and</strong> managed<br />

systems where the performance of one particular function is overexploited.<br />

Overexploitation of user <strong>and</strong> production functions results not only in the depletion of the<br />

resource s<strong>to</strong>ck or flow provided but also in the inability of other functions <strong>to</strong> perform <strong>to</strong> their<br />

full potential. The same applies when the capacity of regulation functions is exceeded, for<br />

example when coastal waters are polluted beyond their waste-assimilation capacity. Both<br />

forms of unsustainable development can inhibit or destroy the working of functions that are<br />

essential <strong>to</strong> the provision of resources that are valued less financially or <strong>to</strong> the maintenance<br />

of coastal <strong>resilience</strong>. This can ultimately result in the degradation of natural systems that<br />

provide protection against the sea, habitat for many species <strong>and</strong> food for many people.<br />

If, owing <strong>to</strong> internal or external stresses on the coastal system, one or more of the desired<br />

functions cannot perform <strong>to</strong> their full potential, conflicts could arise. For example, if<br />

mangroves are polluted beyond their filtering capacity or logged <strong>and</strong> cleared in an unsustainable<br />

manner, this will be at the expense of processes that enable fish <strong>to</strong> breed <strong>and</strong> be caught<br />

in the same area <strong>and</strong> hence of those depending on fisheries (e.g., Gilbert <strong>and</strong> Janssen, 1998).<br />

The maintenance of all functions at a sustainable level would provide higher economic returns<br />

over a longer period of time.<br />

Traditionally, coastal zone management has been described as a process <strong>to</strong> h<strong>and</strong>le conflicts<br />

between the various (potential) users of increasingly scarce coastal resources <strong>and</strong> <strong>to</strong><br />

address current problems that result from stakeholders pursuing their own sec<strong>to</strong>ral interests<br />

(e.g., Cicin-Sain, 1993; Post <strong>and</strong> Lundin, 1996; Vallega, 1996; Cicin-Sain <strong>and</strong> Knecht, 1998). In<br />

other words, coastal zone management has focused strongly on conflicts stemming from the<br />

multiple uses of resources provided by user <strong>and</strong> production functions.<br />

The purpose of the following sections is <strong>to</strong> show that coastal zone management is also<br />

appropriate <strong>to</strong> tackle or anticipate issues that rely on the robust performance of the coastal<br />

system’s regulation functions. Such issues are typically associated with longer-term developments<br />

that mostly occur on a larger, often global scale. These longer-term global developments,<br />

often denoted as “global change”, include demographic trends <strong>and</strong> economic developments,<br />

as well as human interference with global environmental systems such as climate.<br />

Before Section 3 discusses the implications of climate change for coastal zones, the following<br />

sections focus on demographic trends <strong>and</strong> economic developments.<br />

3


2.2. Demographic trends<br />

There is a long his<strong>to</strong>ry of human settlement in coastal zones, but until the twentieth century<br />

the level of disturbance <strong>to</strong> natural processes did not appear <strong>to</strong> be critical. During the twentieth<br />

century, urbanised coastal populations have been growing rapidly around the globe because<br />

of the many economic opportunities <strong>and</strong> environmental amenities that coastal zones<br />

provide (Turner et al., 1996). Low-lying areas near coasts now have the largest concentrations<br />

of people on earth (Small <strong>and</strong> Cohen, 1999). The population in the “near-coastal zone”<br />

(defined as areas both within 100 m elevation <strong>and</strong> 100 km distance of the coast) in 1990 is estimated<br />

at 1.2 billion (thous<strong>and</strong> million) (Nicholls <strong>and</strong> Small, 2002; see also Cohen et al., 1997<br />

<strong>and</strong> Gommes et al., 1998).<br />

Nicholls <strong>and</strong> Small (2002) also showed that most of the near-coastal zone is sparsely<br />

inhabited, with the human population being concentrated in a few specific areas along the<br />

world’s coast. These areas correspond mainly <strong>to</strong> coastal plains in Europe <strong>and</strong> parts of Asia,<br />

<strong>and</strong> <strong>to</strong> a lesser extent <strong>to</strong> densely populated urban areas. Hence, there are wide variations in<br />

coastal populations amongst nations. In many small isl<strong>and</strong> nations, all l<strong>and</strong> suitable for human<br />

habitation is coastal <strong>and</strong> in some large countries, most or all major urban centres are located<br />

near the coast (e.g., Australia). In other countries, such as Mexico, Colombia, Russia <strong>and</strong> Iran,<br />

many larger cities are found further inl<strong>and</strong>, in spite of the countries’ long coastlines.<br />

The United Nations medium projection for population growth suggests that the world’s<br />

population will reach 7.2 billion by the year 2015, 7.9 billion by 2025 <strong>and</strong> 9.3 billion by 2050<br />

(2000: 6.1 billion; UNPD, 2001). Growth rates in individual countries will be largely determined<br />

by the country’s current demographic patterns <strong>and</strong> fertility rate. Age structures in<br />

most developing countries are such that during the coming decades greater numbers of people<br />

will come in<strong>to</strong> their prime reproductive years than in industrialised countries. Furthermore,<br />

fertility rates are generally higher in developing countries, albeit declining. As a result,<br />

all projected population growth until 2050 is expected <strong>to</strong> occur in the developing world<br />

(UNPD, 2001).<br />

Most of the population growth in developing countries will occur in urban settings <strong>and</strong><br />

much of this will be concentrated in coastal zones, as has been the case in most industrialised<br />

countries. It is projected that by 2015 there will be 33 cities with a population of more than<br />

eight million (UNPD, 2001). As shown in Table 2, 21 of these megacities are located in coastal<br />

zones. This table, which is an update of the overviews provided by WCC’93 (1994) <strong>and</strong><br />

Nicholls (1995a), also shows that seventeen of the 21 largest megacities are coastal <strong>and</strong> that<br />

with the exception of Tokyo, New York, Los Angeles, Osaka, Paris <strong>and</strong> Moscow all projected<br />

megacities are situated in developing countries. Continued growth of urban areas can be expected<br />

after 2015, especially in Africa <strong>and</strong> Asia (UNEP, 2002), resulting in the development of<br />

additional coastal megacities <strong>to</strong> those shown in Table 2.<br />

Some care should be taken in interpreting the data presented in Table 2. A certain degree<br />

of subjectivity is inevitable in labelling a megacity as coastal, especially because there<br />

are no straightforward definitions of a coastal zone. São Paolo, for example, is not considered<br />

coastal because it is situated at an elevation of 800 metres above sea level. However, its<br />

proximity <strong>to</strong> the Atlantic Ocean <strong>and</strong> the port of San<strong>to</strong>s has yielded benefits that would not<br />

have been available in places further inl<strong>and</strong> (e.g., as a coffee trading capital). On the other<br />

h<strong>and</strong>, cities like Dhaka, Calcutta <strong>and</strong> Cairo are situated at some distance from the sea, yet in<br />

Table 2 they are considered coastal because of their deltaic setting. Other cities, such as Los<br />

Angeles, Seoul <strong>and</strong> Istanbul, have developed on coasts with steeper gradients <strong>and</strong> hence parts<br />

of their agglomerations extend outside the coastal plain. The predominant criterion <strong>to</strong> classify<br />

a city as coastal in Table 2 is whether it has economic <strong>and</strong> geomorphic characteristics<br />

that are typically or exclusively coastal (e.g., sea port, deltaic or estuarine setting).<br />

Some coastal agglomerations with populations exceeding 8 million may have been omitted<br />

because of the dataset used for Table 2. These include Greater London in the United<br />

Kingdom <strong>and</strong> the Hong Kong-Shenzhen-Guangzhou conurbation in China (Nicholls, 1995a).<br />

More dispersed agglomerations are not considered either, such as the Amsterdam-Brussels<br />

axis in The Netherl<strong>and</strong>s <strong>and</strong> Belgium, the Osaka-Nagoya-Tokyo axis in Japan <strong>and</strong> ‘Megalopolis’<br />

in the United States, which stretches over 600 km from Bos<strong>to</strong>n, MA <strong>to</strong> Washing<strong>to</strong>n, DC <strong>and</strong> has<br />

a collective population approaching 50 million. Such dispersed coastal agglomerations may<br />

also emerge in the developing world, such as from Accra, Ghana <strong>to</strong> Lagos, Nigeria, embracing<br />

parts of four countries.<br />

4


»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

»<br />

Rank Agglomeration Country Population size (million) Expected<br />

growth (%)<br />

1975 2000 2015 2000–2015<br />

1 Tokyo<br />

Japan<br />

19.771 26.444 26.444 0.00<br />

2 Bombay<br />

India<br />

6.856 18.066 26.138 44.68<br />

3 Lagos<br />

Nigeria<br />

3.300 13.427 23.173 72.59<br />

4 Dhaka<br />

Bangladesh<br />

2.172 12.317 21.119 71.46<br />

5 São Paolo<br />

Brazil<br />

10.047 17.755 20.397 14.88<br />

6 Karachi<br />

Pakistan<br />

3.983 11.794 19.211 62.89<br />

7 Mexico City Mexico<br />

11.236 18.131 19.180 5.79<br />

8 New York<br />

United States of America 15.880 16.640 17.432 4.76<br />

9 Jakarta<br />

Indonesia<br />

4.814 11.018 17.256 56.62<br />

10 Calcutta<br />

India<br />

7.888 12.918 17.252 33.55<br />

11 Delhi<br />

India<br />

4.426 11.695 16.808 43.72<br />

12 Metro Manila Philippines<br />

5.000 10.870 14.825 36.38<br />

13 Shanghai<br />

China<br />

11.443 12.887 14.575 13.10<br />

14 Los Angeles United States of America 8.926 13.140 14.080 7.15<br />

15 Buenos Aires Argentina<br />

9.144 12.560 14.076 12.07<br />

16 Cairo<br />

Egypt<br />

6.079 10.552 13.751 30.32<br />

17 Istanbul<br />

Turkey<br />

3.601 9.451 12.492 32.18<br />

18 Beijing<br />

China<br />

8.545 10.839 12.299 13.47<br />

19 Rio de Janeiro Brazil<br />

7.854 10.582 11.905 12.50<br />

20 Osaka<br />

Japan<br />

9.844 11.013 11.013 0.00<br />

21 Tianjin<br />

China<br />

6.160 9.156 10.713 17.01<br />

22 Hyderabad India<br />

2.086 6.842 10.457 52.84<br />

23 Bangkok<br />

Thail<strong>and</strong><br />

3.842 7.281 10.143 39.31<br />

24 Lahore<br />

Pakistan<br />

2.399 6.040 9.961 64.92<br />

25 Seoul<br />

Republic of Korea<br />

6.808 9.888 9.923 0.35<br />

26 Paris<br />

France<br />

8.885 9.624 9.677 0.55<br />

27 Lima<br />

Peru<br />

3.651 7.443 9.388 26.13<br />

28 Kinshasa<br />

Dem. Rep. of the Congo 1.735 5.064 9.366 84.95<br />

29 Moscow<br />

Russian Federation<br />

7.623 9.321 9.353 0.34<br />

30 Madras<br />

India<br />

3.609 6.648 9.145 37.56<br />

31 Chongqing China<br />

2.439 5.312 8.949 68.47<br />

32 Teheran<br />

Islamic Republic of Iran 4.274 7.225 8.709 20.54<br />

33 Bogotá<br />

Colombia<br />

3.036 6.288 8.006 27.32<br />

Total (coastal agglomerations) 150.6 254.1 324.1 27.53<br />

Total (all agglomerations) 217.4 368.2 467.2 26.88<br />

Rank in<br />

1975<br />

1<br />

15<br />

33<br />

43<br />

5<br />

16<br />

4<br />

2<br />

22<br />

12<br />

23<br />

21<br />

3<br />

8<br />

7<br />

20<br />

32<br />

10<br />

13<br />

6<br />

19<br />

45<br />

28<br />

41<br />

16<br />

9<br />

30<br />

n/a<br />

14<br />

31<br />

40<br />

25<br />

35<br />

Rank in<br />

2000<br />

1<br />

3<br />

6<br />

11<br />

4<br />

12<br />

2<br />

5<br />

14<br />

8<br />

13<br />

16<br />

9<br />

7<br />

10<br />

19<br />

22<br />

17<br />

18<br />

15<br />

24<br />

31<br />

27<br />

35<br />

20<br />

21<br />

25<br />

40<br />

23<br />

32<br />

38<br />

28<br />

34<br />

n/a: not available.<br />

Table 2 — The world’s largest cities, with projected populations in 2015 exceeding eight<br />

million. Arrows indicate coastal agglomerations (population data from UNPD, 2001).<br />

Urban populations tend <strong>to</strong> have higher consumption levels than their rural counterparts,<br />

as well as different consumption patterns. The increasing dem<strong>and</strong> for food requires increasing<br />

productivity in fisheries <strong>and</strong> agriculture. However, this is often impeded by the loss of agricultural<br />

l<strong>and</strong> <strong>to</strong> urban expansion <strong>and</strong> the reduction of fisheries potential due <strong>to</strong> habitat loss<br />

<strong>and</strong> pollution of rivers <strong>and</strong> coastal waters from urban <strong>and</strong> industrial waste.<br />

In addition, the large populations in many coastal areas around the world are, <strong>to</strong> a<br />

greater or lesser extent, vulnerable <strong>to</strong> hazardous events associated with natural coastal dynamics<br />

such as s<strong>to</strong>rm surges, floods <strong>and</strong> tsunamis. Human-induced climate change <strong>and</strong> sealevel<br />

rise will further increase this <strong>vulnerability</strong>, as discussed in Section 3.<br />

2.3. Economic values <strong>and</strong> competing dem<strong>and</strong>s<br />

In addition <strong>to</strong> population growth <strong>and</strong> urbanisation, coastal areas are facing unprecedented<br />

pressures on the economic development of their resources (Turner et al., 1996; Post <strong>and</strong><br />

Lundin, 1996). The diversity of functions performed by natural coastal systems supports a variety<br />

of economic activities. As stated in Section 2.1, important economic activities in coastal<br />

zones include <strong>to</strong>urism <strong>and</strong> recreation, exploitation of living <strong>and</strong> non-living resources, industry<br />

<strong>and</strong> commerce, infrastructure development <strong>and</strong> nature conservation. Industry <strong>and</strong> commerce<br />

<strong>and</strong> infrastructure development are closely related <strong>to</strong> demographic developments as discussed<br />

in Section 2.2. This section focuses on <strong>to</strong>urism <strong>and</strong> recreation, fisheries <strong>and</strong> nature conservation.<br />

2.3.1. Tourism <strong>and</strong> recreation<br />

Tourism represents an important <strong>and</strong> growing economic activity in many parts of the world. In<br />

1995, global international <strong>to</strong>urist receipts amounted <strong>to</strong> US$ 400 billion <strong>and</strong> most of this is<br />

coastal (OECD, 1997). Including domestic travel, <strong>to</strong>urism is the world’s largest single industry<br />

5


<strong>and</strong> the potential for growth is still substantial in many places. Estimates indicate that it accounts<br />

for at least 5% of the combined Gross National Products (GNPs) of all countries, although<br />

wide variations exist. For example, in the early 1990s <strong>to</strong>urism was estimated <strong>to</strong> contribute<br />

about 43% of the Caribbean region’s combined GNP (Miller <strong>and</strong> Auyong, 1991).<br />

Since the late 1970s, a range of textbooks <strong>and</strong> assessment reports have discussed the<br />

close interrelationships between <strong>to</strong>urism <strong>and</strong> the environment (e.g., Eding<strong>to</strong>n <strong>and</strong> Eding<strong>to</strong>n,<br />

1986; Briassoulis <strong>and</strong> Van der Straaten, 1992; Nelson et al., 1993). A common conclusion is<br />

that the viability <strong>and</strong> sustainability of the <strong>to</strong>urism sec<strong>to</strong>r is often undermined by (i) a lack of<br />

consideration of the external effects of sec<strong>to</strong>ral activities on environmental <strong>and</strong> amenity values<br />

that underpin <strong>to</strong>urism <strong>and</strong> (ii) the adverse effects of <strong>to</strong>urism <strong>and</strong> <strong>to</strong>urism development itself.<br />

Thus, whilst coastal <strong>to</strong>urism can be threatened by other developments in the coastal<br />

zone, <strong>to</strong>urist activities themselves may also give rise <strong>to</strong> conflicts.<br />

Activities that could jeopardise <strong>to</strong>urism in coastal areas include aquaculture, agriculture,<br />

industry <strong>and</strong> oil <strong>and</strong> gas exploitation. These activities often compete for the same space <strong>and</strong><br />

resources as those desired by <strong>to</strong>urists. Moreover, they may cause damage <strong>to</strong> coastal <strong>and</strong> marine<br />

habitats <strong>and</strong> thereby degrade the basic resource upon which <strong>to</strong>urism depends <strong>to</strong> thrive<br />

<strong>and</strong> grow (e.g., Dulvy et al., 1995). In addition, eutrophication <strong>and</strong> microbial pollution caused<br />

by the disposal of untreated waste in<strong>to</strong> coastal waters can affect the quality of seafood <strong>and</strong><br />

bathing water <strong>and</strong> thus pose risks <strong>to</strong> human health. Gastroenteritis is the most common of the<br />

diseases that can be contracted (Andreadakis, 1997; Pruss, 1998), but cholera <strong>and</strong> typhoid<br />

can also be developed after the consumption of contaminated seafood (Desenclos, 1996).<br />

Environmental degradation caused by <strong>to</strong>urist development is perhaps most clearly observed<br />

in coral settings. A case study by Hawkins <strong>and</strong> Roberts (1994) examined the impacts of<br />

exp<strong>and</strong>ing coastal <strong>to</strong>urism on coral reefs along the Red Sea coast of Egypt. Activities are in<br />

progress <strong>to</strong> support an almost twelvefold increase in the number of visi<strong>to</strong>rs <strong>to</strong> this area over<br />

the period 1990–2005, 55% of which is likely <strong>to</strong> occur in the two popular resorts Hurghada <strong>and</strong><br />

Sharm-el-Sheikh. Present development in Hurghada has already damaged inshore reefs <strong>to</strong> the<br />

extent that what remains is of little interest <strong>to</strong> divers <strong>and</strong> snorkellers, who are now directed<br />

<strong>to</strong> offshore reefs. The massive expansion plans are likely <strong>to</strong> increase current problems of infilling,<br />

sedimentation, eutrophication, overfishing <strong>and</strong> trampling in both Hurghada <strong>and</strong> Sharmel-Sheikh<br />

unless current regulations are enforced <strong>and</strong> restrictions on <strong>to</strong>urism growth are imposed<br />

(Hawkins <strong>and</strong> Roberts, 1994).<br />

One example of <strong>to</strong>urism-related pollution is connected with the use of tributyltin as an<br />

anti-fouling agent in paints that, in spite of increasingly strict legislation <strong>to</strong> control their usage,<br />

are still applied <strong>to</strong> protect vessel hulls, including those of recreational craft. Tributyltin<br />

is amongst the most <strong>to</strong>xic compounds introduced in<strong>to</strong> the marine environment. It causes reproductive<br />

disorders in many gastropod species such as whelks, of which the females develop<br />

male characteristics—a mechanism referred <strong>to</strong> as imposex (Demora <strong>and</strong> Pelletier, 1997). Alternative<br />

anti-fouling paints are currently under development <strong>and</strong> a <strong>to</strong>tal ban on the use of<br />

tributyltin is expected before 2010 (Ten Hallers-Tjabbes, 1997). However, the problem of imposex<br />

will not be solved overnight, since large concentrations of tributyltin have accumulated<br />

in marine sediments, especially in port areas <strong>and</strong> near shipping routes. Meanwhile, key questions<br />

about <strong>to</strong>xicological mechanisms, particularly in mammals, remain unanswered (Demora<br />

<strong>and</strong> Pelletier, 1997).<br />

In summary, <strong>to</strong>urism can be an important source of income for many coastal countries<br />

but great care must be taken in <strong>to</strong>urism planning <strong>and</strong> management <strong>to</strong> avoid “killing the goose<br />

that lays the golden eggs”. In most coastal nations there are increasing dem<strong>and</strong>s for coastal<br />

sites for <strong>to</strong>urism development; policymakers are facing a major challenge in meeting dem<strong>and</strong>s<br />

whilst protecting the quality of the environment (Knight et al., 1997; White et al., 1997). A<br />

bitter lesson learnt by several Mediterranean <strong>and</strong> Asian coastal nations is that once the quality<br />

of the environment declines, <strong>to</strong>urism revenues also drop. Once the reputation of an area<br />

has been damaged, it is extremely difficult <strong>to</strong> convince people <strong>to</strong> return.<br />

2.3.2. Fisheries<br />

About 950 million people worldwide depend on fish as their primary source of protein, whilst<br />

the fishing industry directly or indirectly employs some 200 million people (WRI, 1996).<br />

Amongst the 40 countries that rank highest in per-capita consumption of marine sources of<br />

protein, all but one are developing countries. In 1995, <strong>to</strong>tal global fish production reached a<br />

6


level of 112.9 million metric <strong>to</strong>ns, 75% of which were marine l<strong>and</strong>ings, 6.3% marine aquaculture,<br />

6.4% inl<strong>and</strong> catch <strong>and</strong> 12.2% inl<strong>and</strong> aquaculture. Of <strong>to</strong>tal global fish production, 81.9<br />

million <strong>to</strong>nnes were used for human consumption; the remainder was reduced <strong>to</strong> fishmeal <strong>and</strong><br />

oil (Grainger, 1997).<br />

Three decades ago, Gull<strong>and</strong> (1971) estimated that the maximum sustainable yield for<br />

traditionally exploited marine species would be about 100 million <strong>to</strong>nnes per year. Despite<br />

fisheries also turning <strong>to</strong> non-traditional species, marine capture has been fairly stable between<br />

78 <strong>and</strong> 86 million <strong>to</strong>nnes since 1986 (although an upward trend can be seen as of 1994),<br />

whilst marine aquaculture has increased from 3 <strong>to</strong> more than 7 million <strong>to</strong>nnes over the same<br />

period (Grainger, 1997). Caught but not included in the statistics are species of low commercial<br />

value <strong>and</strong> juvenile species. This unwanted bycatch has been estimated at between 18 <strong>and</strong><br />

39 million <strong>to</strong>nnes per year (Alverson et al., 1994).<br />

An important sign that marine fish are currently being overexploited is the increasing<br />

catch of mature <strong>and</strong> senescent fish <strong>and</strong> the decreasing catch of undeveloped <strong>and</strong> developing<br />

fish. This suggests that fewer fish are being born than are caught (Grainger <strong>and</strong> Garcia, 1996).<br />

This trend reflects the intensification of fisheries since 1950 <strong>and</strong> the increase in the proportion<br />

of marine fish subject <strong>to</strong> declines in productivity. It also underlines the fact that the<br />

growing <strong>to</strong>tal <strong>to</strong>nnage of fish production over this period gives a misleading picture of the<br />

state of world fisheries <strong>and</strong> a false sense of security (Grainger <strong>and</strong> Garcia, 1996). However,<br />

projections by individual nations, especially China, suggest that marine fish l<strong>and</strong>ings will continue<br />

<strong>to</strong> increase significantly. Continued overexploitation <strong>and</strong> declining productivity could<br />

result in dem<strong>and</strong>s for marine fish exceeding annual production, with the consequent effect of<br />

increased prices <strong>and</strong> reduced availability of marine protein <strong>to</strong> many people in developing<br />

countries.<br />

In addition <strong>to</strong> overexploitation, part of the explanation for the stress on marine fish<br />

s<strong>to</strong>cks lies in the loss of coastal fisheries habitats as a result of marine pollution <strong>and</strong> their<br />

conversion <strong>to</strong> aquaculture or other productive l<strong>and</strong>. These habitats form the breeding, nursery<br />

<strong>and</strong> feeding grounds for many fish <strong>and</strong> crustaceans that are currently exploited. In countries<br />

such as Thail<strong>and</strong> <strong>and</strong> the Philippines, it is estimated that more than 70% of the mangrove<br />

area has been cleared <strong>and</strong> replaced by shrimp ponds (WRI, 1996). The socio-economic impacts<br />

of coastal aquaculture developments that require major alteration of coastal ecosystems can<br />

be far-reaching.<br />

Apart from the direct loss of mangrove <strong>and</strong> other valuable coastal systems, coastal aquaculture<br />

can have a number of other adverse consequences on coastal systems. Reported impacts<br />

include l<strong>and</strong> subsidence, acidification of soils <strong>and</strong> estuarine waters, salinisation of<br />

groundwater <strong>and</strong> agricultural l<strong>and</strong>s <strong>and</strong> the subsequent loss of economic <strong>and</strong> environmental<br />

goods <strong>and</strong> services upon which coastal communities rely (WRI, 1996). <strong>Coastal</strong> aquaculture has<br />

led <strong>to</strong> drops in agricultural productivity <strong>and</strong> farm incomes, reduced water supplies, loss of income<br />

from fishing <strong>and</strong> forestry <strong>and</strong> increasing hazard of coastal flooding. One of the more recent<br />

socio-economic problems resulting from large-scale aquaculture development is the<br />

growing frequency of disease outbreaks, including algae blooms known as red tides, which can<br />

cause mass fish kills, resulting in major losses of income <strong>and</strong> threats <strong>to</strong> public health (WRI,<br />

1996).<br />

2.3.3. Nature conservation<br />

As discussed in Section 2.1, coastal zones provide a range of environmental goods <strong>and</strong> services<br />

that are essential <strong>to</strong> human social <strong>and</strong> economic well-being. Although each ecosystem function<br />

provides different goods <strong>and</strong> services, its performance is often strongly connected <strong>to</strong> that<br />

of other functions. Development in coastal zones is traditionally triggered by the many economic<br />

opportunities provided by the user <strong>and</strong> production functions. However, often there has<br />

been little awareness that regulation functions provide the essential conditions for the full<br />

performance of these user <strong>and</strong> production functions. It is often not recognised that there are<br />

limits <strong>to</strong> the extent <strong>to</strong> which natural coastal systems allow for the exploitation of resources<br />

produced in coastal zones or that radical changes <strong>to</strong> these systems may adversely affect the<br />

availability or regeneration of these resources.<br />

One of the main reasons for the ongoing loss of natural habitat due <strong>to</strong> human activities is<br />

the fact that the importance of nature <strong>and</strong> a healthy environment <strong>to</strong> human well-being is not<br />

fully reflected in economic planning <strong>and</strong> decision-making. It is difficult <strong>to</strong> attach a monetary<br />

7


value <strong>to</strong> the performance of regulation functions, unlike <strong>to</strong> that of many user <strong>and</strong> production<br />

functions. As a result, potential consequences of the loss of regulation functions are often not<br />

taken in<strong>to</strong> account when assessing the costs <strong>and</strong> benefits of coastal development, although<br />

such losses could reduce future opportunities for resource development. According <strong>to</strong> De<br />

Groot (1992b), Turner <strong>and</strong> Adger (1996), Costanza et al. (1997) <strong>and</strong> others, the <strong>to</strong>tal economic<br />

value of ecosystems should be better represented in l<strong>and</strong>-use planning <strong>and</strong> decisionmaking<br />

in order <strong>to</strong> achieve the conservation <strong>and</strong> sustainable utilisation of nature <strong>and</strong> its resources.<br />

Turner <strong>and</strong> Adger (1996) distinguish between a number of different types of values of<br />

ecosystems, which <strong>to</strong>gether make up their <strong>to</strong>tal economic value (Figure 1) 1 . An important research<br />

challenge in environmental <strong>and</strong> ecological economics has been <strong>to</strong> express the different<br />

types of values shown in Figure 1 in monetary terms so as <strong>to</strong> compare them with the direct<br />

costs <strong>and</strong> benefits of investment, for example in cost-benefit analysis. Various methods have<br />

been developed <strong>to</strong> assess the monetary value of ecosystems. In principle it is possible <strong>to</strong> express<br />

each type of value in monetary terms, although more robust methods exist for the<br />

valuation of most direct-use values <strong>and</strong>—<strong>to</strong> some extent—indirect-use values. In practice,<br />

however, some important methodological caveats <strong>and</strong> potential biases can affect the results<br />

of valuation studies. It goes beyond the scope of this section <strong>to</strong> elaborate on these caveats<br />

<strong>and</strong> biases. The theory of monetary valuation of ecosystems is discussed in detail in Pearce<br />

<strong>and</strong> Turner (1990), Hanley <strong>and</strong> Spash (1993), Dixon et al. (1994) <strong>and</strong> Carson et al. (1996).<br />

Total Economic Value<br />

Use Values<br />

Non-Use Values<br />

Direct Use Values Indirect Use Values Option Value Existence Values<br />

Outputs Benefits Benefits Benefits<br />

fish;<br />

fuelwood;<br />

recreation;<br />

transport/navigation.<br />

flood control;<br />

s<strong>to</strong>rm protection;<br />

nutrient cycling;<br />

waste assimilation;<br />

sedimentation;<br />

habitat loss reduction;<br />

groundwater protection.<br />

insurance value of<br />

preserving options for<br />

future use.<br />

value derived merely<br />

from knowing a species<br />

or system is conserved;<br />

value of passing on<br />

natural assets intact <strong>to</strong><br />

future generations;<br />

moral resource value<br />

motivations.<br />

Figure 1 — Different types of ecosystem values, including examples for coastal zones (based<br />

on Turner <strong>and</strong> Adger, 1996).<br />

In general, methods <strong>to</strong> express ecosystem functions in monetary terms are based on human<br />

preferences, as these should reflect the value that people attach <strong>to</strong> these functions. A<br />

basic distinction can be made between expressed-preference methods <strong>and</strong> revealed-preference<br />

methods. A later paper in this thesis reports on the application of the contingent valuation<br />

method, which is based on expressed preferences, <strong>and</strong> the travel-cost method (TCM),<br />

which is a revealed-preference method, <strong>to</strong> assess the value of nature-based recreation in a<br />

coastal wetl<strong>and</strong> in Engl<strong>and</strong> (Klein <strong>and</strong> Bateman, 1998).<br />

An increasing number of studies provide monetary estimates of the value of natural<br />

coastal systems, showing that this value is often considerable <strong>and</strong> usually far exceeds the<br />

(short-term) returns from non-sustainable coastal resource use. Published estimates vary from<br />

US$ 1.5 million <strong>to</strong> 13 million per square kilometre, with an average of between US$ 2–5 million/km<br />

2 for OECD countries <strong>and</strong> US$ 1.25 million/km 2 for developing countries (Fankhauser,<br />

1995). Such numbers increase the awareness of the importance of conserving coastal ecosys-<br />

1<br />

Gren et al. (1994) argue that, strictly speaking, the aggregate of all values is less than the <strong>to</strong>tal economic<br />

value, as ecosystems are considered <strong>to</strong> have primary value that is conditional on the existence <strong>and</strong> maintenance<br />

of the healthy system rather than on any one individual human-use component. In other words, the sum<br />

of the <strong>to</strong>tal system value is greater than the sum of the component parts.<br />

8


tems, especially in the light of anticipated global changes. <strong>Coastal</strong> ecosystems that can perform<br />

their regulation functions, such as protection against erosion <strong>and</strong> waste assimilation, <strong>to</strong><br />

their full potential are more robust <strong>and</strong> resilient <strong>to</strong> sea-level rise <strong>and</strong> pollution. In addition,<br />

De Groot (1992b) argues that the costs involved in conserving nature <strong>and</strong> managing protected<br />

areas could be seen as productive capital, providing employment <strong>and</strong> safeguarding opportunities<br />

for other uses <strong>and</strong> benefits.<br />

3. Climate change <strong>and</strong> coastal zones<br />

Human-induced global climate change <strong>and</strong> associated sea-level rise can have major adverse<br />

consequences for coastal ecosystems <strong>and</strong> societies. Human-induced climate change is caused<br />

by the emission of so-called “greenhouse” gases, which trap long-wave radiation in the upper<br />

atmosphere <strong>and</strong> thus raise atmospheric temperatures. Carbon dioxide is the most important<br />

of these gases <strong>and</strong> its atmospheric concentration has exponentially increased since the beginning<br />

of the industrial revolution as a result of fossil fuel combustion <strong>and</strong> l<strong>and</strong>-use change. In<br />

1800, the atmospheric concentration of carbon dioxide was about 280 parts per million<br />

(ppm); <strong>to</strong>day it is about 350 ppm <strong>and</strong> rising. Similar increases have been observed for other<br />

greenhouse gases such as methane <strong>and</strong> nitrous oxide (Hough<strong>to</strong>n et al., 2001).<br />

Projections of future climate change are based on global scenarios of future emissions of<br />

greenhouse gases. These emission scenarios are subject <strong>to</strong> great uncertainty, as they reflect<br />

patterns of economic development, population growth, consumption <strong>and</strong> so on that are not<br />

easy <strong>to</strong> foresee over a 100-year period. A large number of emission scenarios are used <strong>to</strong> account<br />

for this high degree of uncertainty. The most recent emission scenarios, which formed<br />

the basis of the climate projections of the IPCC Third Assessment Report (TAR), were published<br />

in the IPCC Special Report on Emission Scenarios (SRES; Nakićenović et al., 2000) <strong>and</strong><br />

are known as the SRES scenarios.<br />

By 2100, carbon cycle models project atmospheric carbon dioxide concentrations of 540<br />

<strong>to</strong> 970 ppm for the illustrative SRES scenarios, with a range of uncertainty of 490 <strong>to</strong> 1260 ppm<br />

(Hough<strong>to</strong>n et al., 2001). Based on these projections <strong>and</strong> those of other greenhouse gases, the<br />

IPCC TAR projects an increase in globally averaged surface temperature by 1.4 <strong>to</strong> 5.8°C over<br />

the period 1990 <strong>to</strong> 2100. These results are for the full range of 35 SRES scenarios, based on a<br />

number of climate models. The IPCC TAR further states that it is very likely that nearly all<br />

l<strong>and</strong> areas will warm more rapidly than the global average, particularly those at northern high<br />

latitudes in the cold season (Hough<strong>to</strong>n et al., 2001).<br />

3.1. Mechanisms <strong>and</strong> projections of sea-level rise<br />

Based on these projections of future climate, global mean sea level is projected <strong>to</strong> rise by 9<br />

<strong>to</strong> 88 cm between 1990 <strong>and</strong> 2100, with a central value of 48 cm, for the full range of SRES<br />

scenarios. The projected sea-level rise is due primarily <strong>to</strong> the thermal expansion of ocean water<br />

(11 <strong>to</strong> 43 cm), followed by contributions from mountain glaciers (1 <strong>to</strong> 23 cm) <strong>and</strong> ice caps<br />

(–2 <strong>to</strong> 9 cm for Greenl<strong>and</strong> <strong>and</strong> –17 <strong>to</strong> 2 cm for Antarctica) (Church et al., 2001). The projected<br />

central value of 48 cm gives an average rate of sea-level rise of 2.2 <strong>to</strong> 4.4 times the observed<br />

rate over the 20th century. Even with drastic reductions in greenhouse gas emissions, sea<br />

level will continue <strong>to</strong> rise for centuries beyond 2100 because of the long response time of the<br />

global ocean system. An ultimate sea-level rise of 2 <strong>to</strong> 4 metres is possible for atmospheric<br />

carbon dioxide concentrations that are twice <strong>and</strong> four times pre-industrial levels, respectively<br />

(Church et al., 2001).<br />

A major uncertainty is how the global mean sea-level rise will manifest itself on a regional<br />

scale. All models analysed by the IPCC show a strongly non-uniform spatial distribution<br />

of sea-level rise. Some regions show a sea-level rise substantially higher than the global average<br />

(in many cases more than twice the average) <strong>and</strong> others a sea-level fall (Church et al.,<br />

2001). However, the patterns produced by the different models are not similar in detail. This<br />

lack of similarity means that confidence in projections of regional sea-level changes is low,<br />

although it is clear that they are important with respect <strong>to</strong> impacts of sea-level rise on<br />

coastal zones.<br />

9


Many coastal areas experience vertical l<strong>and</strong> movements, which change the level of the<br />

sea in relation <strong>to</strong> that of the l<strong>and</strong>, irrespective of absolute sea-level changes. These movements<br />

are often associated with geological processes, but can also result from the extraction<br />

of water <strong>and</strong> hydrocarbons <strong>and</strong> from the oxidation of peat. Excessive groundwater withdrawal<br />

is a particular problem in many coastal cities, exacerbated by the cities’ rapid growth. As the<br />

water table beneath a city drops, formerly saturated sediments can consolidate irreversibly,<br />

causing the ground surface <strong>to</strong> subside rapidly. This subsidence leads <strong>to</strong> an additional rise in<br />

sea level relative <strong>to</strong> the l<strong>and</strong>.<br />

Subsidence rates can be rapid, locally reaching one metre per decade (Nicholls, 1995a).<br />

For instance, l<strong>and</strong> subsidence around Tianjin was around 5 cm/yr in the late 1980s <strong>and</strong> locally<br />

up <strong>to</strong> 11 cm/yr. In Japan, on the other h<strong>and</strong>, subsidence has been largely controlled by means<br />

of better groundwater management, although large areas in the Osaka-Tokyo conurbation,<br />

home <strong>to</strong> 2 million people, are now beneath high-water levels <strong>and</strong> would be submerged but for<br />

the extensive flood defence systems (Mimura et al., 1994).<br />

In addition <strong>to</strong> sea-level rise resulting from anthropogenic greenhouse gas emissions <strong>and</strong><br />

l<strong>and</strong> subsidence, human activities can also influence sea level directly by changing the<br />

amount of water s<strong>to</strong>red in the ground, in lakes <strong>and</strong> in reservoirs <strong>and</strong> by modifying surface<br />

characteristics affecting runoff <strong>and</strong> evapotranspiration rates. These changes are likely <strong>to</strong> have<br />

led <strong>to</strong> a net change in sea level over the past century, although the direction of this change is<br />

disputed. Based on studies by Sahagian et al. (1994), Gornitz et al. (1997), Sahagian (2000),<br />

Gornitz (2000) <strong>and</strong> Vörösmarty <strong>and</strong> Sahagian (2000), Church et al. (2001) estimated that the<br />

average rate with which these changes contributed <strong>to</strong> global sea-level rise between 1910 <strong>and</strong><br />

1990 was –1.1 <strong>to</strong> 0.4 mm/yr.<br />

Sea-level rise is not the only climate-related effect relevant <strong>to</strong> coastal zones. However,<br />

confidence in model projections of other manifestations of climate change is generally still<br />

low. An important issue is the extent <strong>to</strong> which the frequency, intensity <strong>and</strong> spatial patterns of<br />

extreme events will change. A rise in mean sea level will lead <strong>to</strong> a decrease in the return period<br />

of s<strong>to</strong>rm surges, but it is unclear if the variability of s<strong>to</strong>rm surges itself will change.<br />

Changes that are considered likely over some areas are increases in peak wind intensities <strong>and</strong><br />

mean <strong>and</strong> peak precipitation intensities of tropical cyclones. However, changes in tropical cyclone<br />

location <strong>and</strong> frequency are still uncertain (Hough<strong>to</strong>n et al., 2001).<br />

Changes in wind patterns <strong>and</strong> precipitation at higher latitudes also remain uncertain on<br />

local <strong>and</strong> regional scales. To date, no reliable projections exist of coastal s<strong>to</strong>rm characteristics<br />

<strong>and</strong> tidal range on scales useful for coastal impact analysis. The effect of s<strong>to</strong>rm surges<br />

can be compounded by increased precipitation on l<strong>and</strong> <strong>and</strong> associated increased river runoff.<br />

Projections of runoff from selected rivers are becoming available, allowing their consideration<br />

in future analyses.<br />

The IPCC TAR states that recent trends for El Niño are projected <strong>to</strong> continue in many<br />

models (Hough<strong>to</strong>n et al., 2001). These trends include a stronger warming of the eastern<br />

tropical Pacific relative <strong>to</strong> the western tropical Pacific, with a corresponding eastward shift of<br />

precipitation. Changes in El Niño patterns are likely <strong>to</strong> be relevant in a number of coastal locations<br />

but they are still <strong>to</strong>o uncertain <strong>to</strong> be considered in coastal impact studies.<br />

In summary, whilst it is recognised that impacts of climate change on coastal zones are<br />

prompted by more than sea-level rise alone, consideration of other fac<strong>to</strong>rs is constrained by<br />

the uncertainties surrounding them, particularly on local <strong>and</strong> regional scales, with the<br />

emerging exception of river runoff. For this reason, the remainder of this chapter <strong>and</strong> indeed<br />

this thesis will concentrate on sea-level rise as the predominant focus in the analysis of <strong>vulnerability</strong>,<br />

<strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong> in coastal zones.<br />

3.2. Potential impacts of sea-level rise on coastal zones<br />

For coastal areas, it is not the global, absolute sea level that matters but the locally observed,<br />

relative sea level, which takes in<strong>to</strong> account regional sea-level variations <strong>and</strong> vertical<br />

movements of the l<strong>and</strong>. Some parts of the world, such as most of Sc<strong>and</strong>inavia, experience<br />

l<strong>and</strong> uplift at such a high rate that projected absolute sea-level rise will be completely offset<br />

<strong>and</strong> relative sea level will continue <strong>to</strong> fall, albeit at a lower rate. Other areas, such as often<br />

densely populated deltas, are characterised by a strong downward movement of the l<strong>and</strong>,<br />

which will add <strong>to</strong> absolute sea-level rise (Emery <strong>and</strong> Aubrey, 1991; Gröger <strong>and</strong> Plag, 1993).<br />

10


Irrespective of the primary cause of sea-level rise (climate change, natural or human-induced<br />

subsidence, dynamic ocean effects), exposed natural coastal systems can be affected<br />

in a variety of ways. From a societal perspective, the six most important biogeophysical effects<br />

are (Klein <strong>and</strong> Nicholls, 1998):<br />

Increasing flood frequency probabilities;<br />

Erosion;<br />

Inundation;<br />

Rising water tables;<br />

Saltwater intrusion;<br />

Biological effects.<br />

These biogeophysical effects will have consequent effects on ecosystems <strong>and</strong> eventually<br />

affect socio-economic systems in the coastal zone. However, owing <strong>to</strong> the great diversity <strong>and</strong><br />

variation of natural coastal systems <strong>and</strong> <strong>to</strong> the local <strong>and</strong> regional differences in relative sealevel<br />

rise, the occurrence of <strong>and</strong> response <strong>to</strong> these effects will not be uniform around the<br />

globe. <strong>Coastal</strong> environments particularly at risk include tidal deltas <strong>and</strong> low-lying coastal<br />

plains, s<strong>and</strong>y beaches <strong>and</strong> barrier isl<strong>and</strong>s, coastal wetl<strong>and</strong>s, estuaries <strong>and</strong> lagoons <strong>and</strong> coral<br />

reefs <strong>and</strong> a<strong>to</strong>lls (Bijlsma et al., 1996). Increased coastal flooding is expected <strong>to</strong> be most severe<br />

in South <strong>and</strong> Southeast Asia, Africa, the southern Mediterranean coasts, the Caribbean<br />

<strong>and</strong> most isl<strong>and</strong>s in the Indian <strong>and</strong> Pacific Oceans (Watson et al., 1998; Nicholls et al., 1999).<br />

The effects of climate change <strong>and</strong> associated sea-level rise threaten economic sec<strong>to</strong>rs <strong>to</strong><br />

a varying extent. The potential socio-economic impacts of sea-level rise can be categorised as<br />

follows (Klein <strong>and</strong> Nicholls, 1998):<br />

Direct loss of economic, ecological, cultural <strong>and</strong> subsistence values through loss of l<strong>and</strong>,<br />

infrastructure <strong>and</strong> coastal habitats;<br />

Increased flood risk <strong>to</strong> people, l<strong>and</strong> <strong>and</strong> infrastructure <strong>and</strong> economic, ecological, cultural<br />

<strong>and</strong> subsistence values;<br />

Other impacts related <strong>to</strong> changes in water management, salinity <strong>and</strong> biological activity.<br />

Table 3 lists the most important socio-economic sec<strong>to</strong>rs in coastal zones <strong>and</strong> indicates<br />

from which of the aforementioned biogeophysical effects of climate change they are expected<br />

<strong>to</strong> suffer direct impacts. Indirect impacts, for example impacts on human health resulting<br />

from deteriorating water quality, are also likely <strong>to</strong> be important <strong>to</strong> many sec<strong>to</strong>rs, but<br />

these are not shown in Table 3.<br />

Sec<strong>to</strong>r<br />

Biogeophysical effect<br />

Flood<br />

Water Saltwater Biological<br />

Erosion Inundation<br />

frequency<br />

table rise intrusion effects<br />

Water resources <br />

Agriculture <br />

Human health <br />

Fisheries <br />

Tourism <br />

Human settlements <br />

Table 3 — Qualitative overview of direct socio-economic impacts of climate change on a<br />

number of sec<strong>to</strong>rs in coastal zones (Klein <strong>and</strong> Nicholls, 1998).<br />

Changes in extreme events, whilst still uncertain, can have important consequences for<br />

coastal zones. For example, cyclones in the Bay of Bengal <strong>and</strong> hurricanes in the Caribbean<br />

have already caused serious economic disruption, damage <strong>to</strong> infrastructure <strong>and</strong> loss of human<br />

life, independent of global climate change. As stated, however, the mechanisms that determine<br />

the occurrence of such events, as well as their patterns, are poorly unders<strong>to</strong>od.<br />

In the 1990s, a large <strong>and</strong> concerted effort was made <strong>to</strong> assess the implications of sealevel<br />

rise on coastal countries. Studies have been carried out on local, national <strong>and</strong> regional<br />

scales (e.g., Jeftić et al., 1992, 1996; Tooley <strong>and</strong> Jelgersma, 1992; Ehler, 1993; McLean <strong>and</strong><br />

11


Mimura, 1993; Maul, 1993; O’Callahan, 1994; Nicholls <strong>and</strong> Leatherman, 1995; Lenhart et al.,<br />

1996; Leatherman, 1997; Watson et al., 1998; Nicholls <strong>and</strong> Mimura, 1998; De la Vega-Leinert<br />

et al., 2000a, 2000b; Brochier <strong>and</strong> Ramieri, 2001; Mimura <strong>and</strong> Yokoki, 2001). In the IPCC Second<br />

Assessment Report, Bijlsma et al. (1996) emphasised that most coastal areas are vulnerable<br />

<strong>to</strong> the adverse consequences of sea-level rise <strong>to</strong> some degree <strong>and</strong> that some form of <strong>adaptation</strong><br />

will be necessary. However, the studies show considerable variation in possible impacts.<br />

This is illustrated by Table 4, which provides an overview of quantitative results of national<br />

coastal <strong>vulnerability</strong> studies.<br />

Antigua<br />

Argentina<br />

Bangladesh<br />

Belize<br />

Benin<br />

China<br />

Egypt<br />

Guyana<br />

Japan<br />

Kiribati<br />

Malaysia<br />

Marshall Isl.<br />

Mauritius<br />

Netherl<strong>and</strong>s<br />

Nigeria<br />

Pol<strong>and</strong><br />

Senegal<br />

St. Kitts-Nevis<br />

Tonga<br />

Uruguay<br />

United States<br />

Venezuela<br />

Without Measures<br />

With Measures<br />

People affected People<br />

at risk<br />

Capital value<br />

at loss<br />

L<strong>and</strong> at loss Wetl<strong>and</strong><br />

at loss<br />

People<br />

at risk<br />

Protection<br />

costs (100 yr)<br />

Protection<br />

costs/year<br />

# People<br />

(1,000)<br />

% Total # People US$<br />

(million)<br />

% GNP km 2 % Total km 2 # People US$ (million) % GNP<br />

38 50 1,900<br />

3 7,700<br />

76 0.32<br />

5,600 6<br />

1,100<br />

1,800/3,300 0.02/0.04<br />

71,000 60<br />

5,800<br />

>1,000 >0.06<br />

70 35<br />

1,350 25<br />

126 12<br />

85<br />

>430 >0.41<br />

72,000 7<br />

4,700 9 30,000 59,272 204<br />

120,000 13,133 0.45<br />

600 80 60,000 4,000 1115<br />

500<br />

200 0.26<br />

15,400 15<br />

807,000 72<br />

>200,000 >0.15<br />

9 100<br />

2 8<br />

3 0.10<br />

40<br />

6<br />

10,000<br />

3,200<br />

235<br />

>110<br />

30<br />

13<br />

100<br />

1<br />

67<br />

4<br />

1<br />

>1<br />

47<br />

380<br />

12,286<br />

1,400/1,800<br />

1,500<br />

1,000/2,200<br />

53<br />

1,000/3,800<br />

>143,000<br />

1,700/2,600<br />

56 378,961/<br />

385,761<br />

>7.04<br />

0.05<br />

0.04/0.05<br />

0.02<br />

0.21/0.40<br />

2.65<br />

0.12/0.46<br />

>0.03<br />

0.03/0.04<br />

Table 4 — Quantitative results of national coastal <strong>vulnerability</strong> studies (Nicholls, 1995b).<br />

The global <strong>vulnerability</strong> assessments carried out by Hoozemans et al. (1993) <strong>and</strong> Baarse<br />

(1995) suggest that some 189 million people presently live below the once-per-1000-years<br />

s<strong>to</strong>rm-surge level (the “hazard zone”). They estimate that, under present conditions, an average<br />

of 46 million people per year experience s<strong>to</strong>rm-surge flooding. This number would double<br />

if sea level rises 50 cm (92 million people/year) <strong>and</strong> almost triple if it rises one metre<br />

(118 million people/year). Between 86% <strong>and</strong> 92% of these people would experience flooding<br />

even more than once a year (Baarse, 1995). These projections do not take in<strong>to</strong> account any<br />

further population growth, changes in s<strong>to</strong>rm frequencies <strong>and</strong> intensities or adaptive responses.<br />

In a recent update, Nicholls (2002) estimates the current number of people living in the<br />

hazard zone at 197 million <strong>and</strong> the average annual number of people flooded at 10 million.<br />

When population growth is considered, the former figure would increase <strong>to</strong> 399–598 million in<br />

2100 in the absence of sea-level rise <strong>and</strong> <strong>to</strong> 503–755 million in 2100 when a high sea-level rise<br />

scenario is assumed (96 cm in 2100). Assuming no upgrade in protection levels <strong>and</strong> the same<br />

high sea-level rise scenario, the average annual number of people flooded would be 326–510<br />

million in 2100, of whom 309–484 million would be flooded more than once per year. Assuming<br />

that protection levels increase with growing national income reduces the number <strong>to</strong> 211–<br />

337 million in 2100, of whom 195–311 million would experience flooding more than once per<br />

year.<br />

3.3. Methodologies for assessing impacts of sea-level rise<br />

Most studies that provided the source data for Table 4 applied the Common Methodology<br />

for Assessing Vulnerability <strong>to</strong> Sea Level Rise, which was developed by the former <strong>Coastal</strong> Zone<br />

Management Subgroup of the IPCC (IPCC CZMS, 1992). The Common Methodology was meant<br />

<strong>to</strong> assist countries in making first-order assessments of potential coastal impacts of <strong>and</strong> adap-<br />

12


tations <strong>to</strong> sea-level rise (for a more elaborate discussion see Klein <strong>and</strong> Nicholls, 1999). It was<br />

used as the basis of assessments in at least 46 countries. The assessments aimed at identifying<br />

populations <strong>and</strong> resources at risk <strong>and</strong> the costs <strong>and</strong> feasibility of possible responses <strong>to</strong> adverse<br />

impacts. Quantitative results were produced in 22 country case studies (shown in Table<br />

4) <strong>and</strong> eight subnational studies (Nicholls, 1995b).<br />

It is important that the quantitative results in Table 4 are interpreted as being indicative<br />

only. Studies using the Common Methodology were meant <strong>to</strong> serve as prepara<strong>to</strong>ry assessments,<br />

identifying priority regions <strong>and</strong> priority sec<strong>to</strong>rs <strong>and</strong> providing an initial screening of<br />

possible measures. Scientific uncertainties are compounded by uncertainties surrounding future<br />

socio-economic developments <strong>and</strong> by the fact that the damage cost estimates are sensitive<br />

<strong>to</strong> the use of discount rates. In addition, the impacts shown in Table 4 have been assessed<br />

assuming a one-metre rise in sea level by 2100, whereas the latest scientific information,<br />

as presented in Section 3.1, suggests a lower global mean sea-level rise. On the other<br />

h<strong>and</strong>, one can argue that the results may well represent underestimates, because impacts on<br />

non-market values have not been assessed <strong>and</strong> because the studies assume sea-level rise <strong>to</strong><br />

be a gradual process. As shown by West <strong>and</strong> Dowlatabadi (1999) <strong>and</strong> West et al. (2001), superimposing<br />

current s<strong>to</strong>rm-surge variability on a gradual rise in sea level can lead <strong>to</strong> estimates<br />

of damage costs that are an order of magnitude different from those based only on<br />

gradual changes, where perfect foresight is assumed (cf. Yohe et al., 1996; Yohe <strong>and</strong> Neumann,<br />

1997).<br />

As noted by Klein <strong>and</strong> Nicholls (1999), many studies that used the Common Methodology<br />

faced a lack of the accurate <strong>and</strong> complete data necessary for impact <strong>and</strong> <strong>adaptation</strong> assessment.<br />

This has limited its applicability, which has led <strong>to</strong> the development of alternative assessment<br />

methodologies (e.g., Kay <strong>and</strong> Hay, 1993; Gornitz et al., 1994). Nonetheless, no<br />

other methodology has been applied as widely <strong>and</strong> evaluated as thoroughly as the Common<br />

Methodology (e.g., Kay et al., 1996). Thus, it has contributed <strong>to</strong> underst<strong>and</strong>ing the consequences<br />

of sea-level rise <strong>and</strong> encouraged long-term thinking about coastal zones. It also became<br />

a model for assessing impacts <strong>and</strong> <strong>adaptation</strong> in non-coastal systems.<br />

In 1994, the IPCC published its Technical Guidelines for Assessing Climate Change Impacts<br />

<strong>and</strong> Adaptations (Carter et al., 1994), which provide generic guidance <strong>to</strong> countries that<br />

wish <strong>to</strong> assess their <strong>vulnerability</strong> <strong>to</strong> climate change. For a range of socio-economic <strong>and</strong> physiographic<br />

systems, the United Nations Environment Programme (UNEP) H<strong>and</strong>book on Methods<br />

for Climate Change Impact Assessments <strong>and</strong> Adaptation Strategies (Feenstra et al., 1998) then<br />

offers a detailed elaboration of the IPCC Technical Guidelines, including for coastal zones<br />

(Klein <strong>and</strong> Nicholls, 1998). The similarities <strong>and</strong> differences between the IPCC Common Methodology<br />

<strong>and</strong> the IPCC Technical Guidelines are discussed in Klein <strong>and</strong> Nicholls (1999). Klein et<br />

al. (1999) evaluate the guidance on <strong>adaptation</strong> assessment provided by the IPCC Technical<br />

Guidelines.<br />

Adaptation can play an important part in reducing the potential impacts of climate<br />

change in coastal zones. In both the IPCC Common Methodology <strong>and</strong> the IPCC Technical<br />

Guidelines, the guidance provided for <strong>adaptation</strong> assessment is limited. Rather than evaluating<br />

a range of options using some formal decision-making framework (e.g., cost-benefit or<br />

multi-criteria analysis), studies tended <strong>to</strong> consider only a protection versus a do-nothing scenario.<br />

However, there are many different ways <strong>and</strong> options <strong>to</strong> respond <strong>to</strong> the impacts of sealevel<br />

rise (see Section 4.3).<br />

In fact, many of the early studies used a so-called “dumb farmer” scenario 2 : they assumed<br />

that present-day behaviour <strong>and</strong> activities would continue unchanged in the future, irrespective<br />

of how they might be affected by climate change. By ignoring any <strong>adaptation</strong>,<br />

these studies did not distinguish between potential <strong>and</strong> residual impacts <strong>and</strong> thus their damage-cost<br />

values represent serious overestimates. On the other h<strong>and</strong>, the studies, which are<br />

not unique <strong>to</strong> agriculture, served <strong>to</strong> generate awareness of the potential magnitude of impacts<br />

<strong>and</strong> the need for anticipa<strong>to</strong>ry <strong>adaptation</strong>.<br />

2<br />

The dumb farmer is a metaphor for any impacted economic agent that does not anticipate climate change or<br />

act upon its manifestation. Instead, it continues <strong>to</strong> act as if nothing has changed. By not responding <strong>to</strong> changing<br />

circumstances, the agent reduces its profitability or fails <strong>to</strong> take advantage of emerging opportunities. It thus<br />

incurs larger damages than would have been the case had some <strong>adaptation</strong> taken place. The clairvoyant<br />

farmer, on the other h<strong>and</strong>, has perfect knowledge <strong>and</strong> foresight <strong>and</strong> is able <strong>to</strong> minimise damages or maximise<br />

benefits. As always, reality will be somewhere in between.<br />

13


Current methodological work focuses more strongly on the role of <strong>adaptation</strong> <strong>and</strong> adaptive<br />

capacity in determining <strong>vulnerability</strong> <strong>to</strong> climate change. Methods for so-called secondgeneration<br />

<strong>vulnerability</strong> assessment are being developed, building on the insights developed<br />

in the past ten years. An introduction <strong>to</strong> second-generation <strong>vulnerability</strong> assessment is given<br />

in the synthesis chapter of this thesis.<br />

4. Vulnerability, <strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong>: exploring the concepts<br />

Vulnerability, <strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong> are the three basic concepts around which the research<br />

that forms the basis of this thesis has been organised. This thesis deals with the <strong>vulnerability</strong>,<br />

<strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong> of coastal zones <strong>to</strong> climate change. However, the three<br />

terms are not used exclusively in this context. In fact, their usage in relation <strong>to</strong> climate<br />

change is relatively new compared <strong>to</strong> that in disciplines such as natural hazard assessment,<br />

ecology <strong>and</strong> economics. It is therefore useful for the analysis of coastal <strong>vulnerability</strong>, <strong>resilience</strong><br />

<strong>and</strong> <strong>adaptation</strong> <strong>to</strong> climate change <strong>to</strong> underst<strong>and</strong> how the concepts have been defined<br />

<strong>and</strong> used in other fields. This section presents an overview.<br />

4.1. Vulnerability<br />

Vulnerability is an important concept in the United Nations Framework Convention on Climate<br />

Change (UNFCCC), which is the single most important document on climate change for both<br />

scientists <strong>and</strong> policymakers. The UNFCCC was one of the products of the United Nations Conference<br />

on Environment <strong>and</strong> Development (UNCED), which was held in Rio de Janeiro, Brazil,<br />

in June 1992. However, in referring <strong>to</strong> countries that are “particularly vulnerable <strong>to</strong> the adverse<br />

effects of climate change” (Article 4.4) it does not provide a definition of <strong>vulnerability</strong><br />

or any other indication as <strong>to</strong> how particularly vulnerable countries may be identified. The<br />

IPCC, as the main international scientific body on climate change, <strong>to</strong>ok up the challenge <strong>to</strong><br />

develop a methodology for assessing <strong>vulnerability</strong> <strong>to</strong> climate change.<br />

As mentioned in Section 3.3, the then <strong>Coastal</strong> Zone Management Subgroup of the IPCC<br />

developed a Common Methodology for Assessing Vulnerability <strong>to</strong> Sea Level Rise (IPCC CZMS,<br />

1992), which formed the inspiration for the generic IPCC Technical Guidelines for Assessing<br />

Climate Change Impacts <strong>and</strong> Adaptations (Carter et al., 1994). The idea behind a common<br />

methodology <strong>and</strong> technical guidelines was that if all countries were <strong>to</strong> conduct similar assessments,<br />

their relative vulnerabilities <strong>to</strong> climate change could be compared <strong>and</strong> it would<br />

become clear which of them are particularly vulnerable.<br />

In addition, knowledge of <strong>vulnerability</strong> would enable coastal scientists <strong>and</strong> policymakers<br />

<strong>to</strong> anticipate impacts that could result from climate change <strong>and</strong> sea-level rise. It could thus<br />

help <strong>to</strong> prioritise management efforts that need <strong>to</strong> be undertaken <strong>to</strong> minimise risks or reduce<br />

possible consequences. Vulnerability assessments were then seen as contributing <strong>to</strong> integrated<br />

coastal zone management programmes, which are based on an awareness of the types<br />

<strong>and</strong> magnitude of problems that different coastal areas may have <strong>to</strong> face, as well as of possible,<br />

alternative solutions (WCC’93, 1994; Bijlsma et al., 1996; see Section 5).<br />

IPCC CZMS (1992) defines the <strong>vulnerability</strong> of coastal zones as “the degree of incapability<br />

<strong>to</strong> cope with the consequences of climate change <strong>and</strong> accelerated sea-level rise”. This definition<br />

implies that <strong>vulnerability</strong> is a composite measure of anticipated impacts of climate<br />

change <strong>and</strong> the extent <strong>to</strong> which systems can adapt <strong>to</strong> these impacts (see Section 4.3). It is<br />

consistent with earlier conceptual work on <strong>vulnerability</strong>, which usually presents <strong>vulnerability</strong><br />

as a reverse function of a system’s ability <strong>to</strong> cope with stress <strong>and</strong> shock (e.g., Timmerman,<br />

1981; Smith, 1992).<br />

The chapter “<strong>Coastal</strong> Zones <strong>and</strong> Small Isl<strong>and</strong>s” of the IPCC Second Assessment Report<br />

(Bijlsma et al., 1996) also emphasises that <strong>vulnerability</strong> is a multi-dimensional concept, encompassing<br />

biogeophysical, socio-economic <strong>and</strong> political fac<strong>to</strong>rs. It can then be argued that<br />

the results presented in Table 4 only show part of the picture of <strong>vulnerability</strong>, as the “incapability<br />

<strong>to</strong> cope” is measured only as the financial cost of coastal protection relative <strong>to</strong> a<br />

country’s economy. Whether coastal protection is the most appropriate <strong>adaptation</strong> strategy in<br />

all cases is not considered, nor are any other indica<strong>to</strong>rs of “incapability <strong>to</strong> cope” analysed.<br />

Increasing recognition of this methodological flaw has led <strong>to</strong> a number of alternative ap-<br />

14


proaches (e.g., Harvey et al., 1999), culminating in the emergence of the concept of “adaptive<br />

capacity” (see Sections 4.3 <strong>and</strong> 6.2).<br />

Before climate change emerged as an academic focus in coastal research, <strong>vulnerability</strong><br />

as such was not an important concept. Traditional research in coastal zones is conducted<br />

mainly by geologists, ecologists <strong>and</strong> engineers, roughly as follows:<br />

Geologists study coastal sedimentation patterns <strong>and</strong> the consequent dynamic processes<br />

of erosion <strong>and</strong> accretion over different spatial <strong>and</strong> temporal scales;<br />

Ecologists study the occurrence, diversity <strong>and</strong> functioning of coastal flora <strong>and</strong> fauna from<br />

the species <strong>to</strong> the ecosystem level;<br />

Engineers take a risk-based approach, assessing the probability of occurrence of s<strong>to</strong>rm<br />

surges <strong>and</strong> other extreme events that could jeopardise the integrity of the coast <strong>and</strong> the<br />

safety of coastal communities.<br />

Thus, the coastal scientific community in general <strong>and</strong> the IPCC <strong>Coastal</strong> Zone Management<br />

Subgroup in particular were not hindered by knowledge of existing conceptualisations of <strong>vulnerability</strong><br />

when defining coastal <strong>vulnerability</strong> <strong>to</strong> climate change <strong>and</strong> developing a methodology<br />

for its assessment. The approach taken for coastal zones was seen as a model for other<br />

systems within the IPCC <strong>and</strong> <strong>vulnerability</strong> <strong>to</strong> climate change is now generally interpreted <strong>and</strong><br />

assessed in this manner.<br />

This approach <strong>to</strong> <strong>vulnerability</strong> assessment has been intuitively attractive <strong>to</strong> the climate<br />

change community, which is strongly model-orientated. Based on scenarios <strong>and</strong> models of<br />

greenhouse gas emissions <strong>and</strong> climate change, sec<strong>to</strong>ral models (e.g., hydrological, ecological,<br />

agricultural <strong>and</strong> disease vec<strong>to</strong>r models) can be applied <strong>to</strong> project potential biophysical impacts,<br />

which can then be evaluated in terms of damage <strong>to</strong> a country’s or sec<strong>to</strong>r’s economy.<br />

The role of <strong>adaptation</strong> in such assessments is briefly discussed in Section 4.3.<br />

Many in the climate change community were unaware of the long his<strong>to</strong>ry of <strong>vulnerability</strong><br />

assessment in other disciplines, particularly in the fields of food security <strong>and</strong> natural hazard<br />

reduction. In these fields, <strong>vulnerability</strong> is also interpreted in terms of potential harm <strong>and</strong> capacity<br />

<strong>to</strong> cope, but there is a crucial difference in that the geographical scale of climate<br />

change <strong>vulnerability</strong> studies is considerably coarser than that of food security <strong>and</strong> natural<br />

hazard studies. The latter type of studies tend <strong>to</strong> focus in more depth on particular groups<br />

<strong>and</strong> communities within a society. In so doing, they take a radically different approach <strong>to</strong> <strong>vulnerability</strong><br />

assessment.<br />

As stated, climate change <strong>vulnerability</strong> studies typically take model results as an input <strong>to</strong><br />

assess potential consequences of climate change on a country, sec<strong>to</strong>r or physiographic system.<br />

The resolution of climate <strong>and</strong> sec<strong>to</strong>ral models <strong>and</strong> the uncertainty surrounding their<br />

output inhibit meaningful assessments of impacts <strong>and</strong> <strong>adaptation</strong> needs at a local level. In<br />

contrast <strong>to</strong> this somewhat mechanistic <strong>to</strong>p-down approach is the bot<strong>to</strong>m-up approach of <strong>vulnerability</strong><br />

analysis in food security <strong>and</strong> natural hazard studies. This approach is typically<br />

place-based <strong>and</strong> cognisant of the rich variety of social, cultural, economic, institutional <strong>and</strong><br />

other fac<strong>to</strong>rs that define <strong>vulnerability</strong> <strong>to</strong> famine or natural hazards. It does not rely on global<br />

or regional models <strong>to</strong> inform the analysis; instead the major source of information is the vulnerable<br />

community itself.<br />

It is clear that climate change, food security <strong>and</strong> natural hazards are related. In some areas<br />

climate change could threaten food security, whilst in many areas the frequency <strong>and</strong> intensity<br />

of weather-related natural hazards are likely <strong>to</strong> increase. There is growing recognition<br />

of the fact that many climate change <strong>vulnerability</strong> studies, whilst useful in alerting policymakers<br />

<strong>to</strong> the potential consequences of climate change, have limited usefulness in providing<br />

guidance on <strong>adaptation</strong> at local scales. These studies do not consider the complex nature of<br />

<strong>vulnerability</strong> reflected in the definition of Downing et al. (1996):<br />

Vulnerability is a complex <strong>and</strong> multidimensional social space defined by the determinate political, economic<br />

<strong>and</strong> institutional capabilities of people in specific places at specific times.<br />

In other words, <strong>vulnerability</strong> is a prior condition, bound up with the social <strong>and</strong> economic<br />

situation of households <strong>and</strong> communities, <strong>and</strong> although external physical fac<strong>to</strong>rs (such as climate<br />

change) play a role in raising <strong>vulnerability</strong>, they are not preconditions or sole causes<br />

(Adger <strong>and</strong> Kelly, 1999).<br />

15


In the climate change community, a process has started <strong>to</strong> complement the <strong>to</strong>p-down<br />

studies, which consider climate change as the sole driver of <strong>vulnerability</strong>, with more placebased<br />

studies, which aim <strong>to</strong> analyse the non-climatic context influencing <strong>vulnerability</strong>, so as<br />

<strong>to</strong> integrate the need for <strong>adaptation</strong> with other, more imminent priorities of vulnerable<br />

communities. Some observations on the role of the international climate community in this<br />

process are presented in the final chapter of this thesis.<br />

4.2. Resilience<br />

The extent <strong>to</strong> which a coastal system is affected by sea-level rise will strongly depend on its<br />

<strong>resilience</strong> <strong>to</strong> changes. Non-climate stresses may already have adversely affected the coastal<br />

system’s <strong>resilience</strong> <strong>and</strong> thereby its ability <strong>to</strong> cope with additional pressures (see Section 2). In<br />

addition, at those places where the coast has been developed <strong>and</strong> protected by hard infrastructure,<br />

l<strong>and</strong>ward migration of coastal ecosystems such as wetl<strong>and</strong>s is obstructed (Bijlsma<br />

et al., 1996).<br />

<strong>Coastal</strong> <strong>resilience</strong> is the <strong>to</strong>pic of a later paper in this thesis (Klein et al., 1998). This paper<br />

distinguishes between morphological, ecological <strong>and</strong> socio-economic <strong>resilience</strong> <strong>and</strong> was<br />

part of a special issue of the Geographical Journal devoted entirely <strong>to</strong> <strong>resilience</strong> in coastal<br />

zones (Nicholls <strong>and</strong> Branson, 1998). The aim of this section is <strong>to</strong> provide a brief overview of<br />

the literature on <strong>resilience</strong> in other disciplines, focusing in particular on ecosystem <strong>resilience</strong><br />

<strong>and</strong> social <strong>resilience</strong>. This section is based on Klein et al. (2002).<br />

The Oxford English Dictionary defines <strong>resilience</strong> as (i) the act of rebounding or springing<br />

back <strong>and</strong> (ii) elasticity. The origin of the word is in Latin, where resilio means <strong>to</strong> jump back.<br />

In a purely mechanical sense, the <strong>resilience</strong> of a material is the quality of being able <strong>to</strong> s<strong>to</strong>re<br />

strain energy <strong>and</strong> deflect elastically under a load without breaking or being deformed (Gordon,<br />

1978). However, since the 1970s the concept has also been used in a more metaphorical<br />

sense <strong>to</strong> describe systems that undergo stress <strong>and</strong> have the ability <strong>to</strong> recover.<br />

Holling (1973) coins the term <strong>resilience</strong> for ecosystems as a measure of the ability of<br />

these systems <strong>to</strong> absorb changes <strong>and</strong> still persist. As such, it determines the persistence of<br />

relationships within an ecosystem. This is contrasted with stability, which is defined by Holling<br />

(1973) as the ability of a system <strong>to</strong> return <strong>to</strong> a state of equilibrium after a temporary disturbance.<br />

Thus, a very stable system would not fluctuate greatly but return <strong>to</strong> normal quickly,<br />

whilst a highly resilient system may be quite unstable, in that it may undergo significant<br />

fluctuation (H<strong>and</strong>mer <strong>and</strong> Dovers, 1996).<br />

Since the seminal work by Holling (1973), <strong>resilience</strong> has become an issue of intense conceptual<br />

debate amongst ecologists. The literature provides many perspectives <strong>and</strong> interpretations<br />

of ecological <strong>resilience</strong> <strong>and</strong>, in spite of thirty years of debate, there appears <strong>to</strong> be no<br />

consensus on how the concept can be made operational or even how it should be defined. Alternative<br />

definitions have been provided, focusing on different system properties. For example,<br />

Pimm (1984) defines <strong>resilience</strong> as the speed with which a system returns <strong>to</strong> its original<br />

state following a perturbation.<br />

Other ecologists question the core assumption that underpins the concept of <strong>resilience</strong>,<br />

namely that ecosystems exist in an equilibrium state <strong>to</strong> which they can return after experiencing<br />

a given level of disturbance. They argue that ecosystems are dynamic <strong>and</strong> evolve continuously<br />

in response <strong>to</strong> external influences taking place on a range of different time scales.<br />

Attempts by ecosystem managers at maintaining some equilibrium state will therefore be<br />

bound <strong>to</strong> fail.<br />

In spite of the relative lack of specificity with which <strong>resilience</strong> has been defined in ecology<br />

(or perhaps as a result of it), the concept has also gained ground in social science, where<br />

it is applied <strong>to</strong> describe the behavioural response of communities, institutions <strong>and</strong> economies.<br />

Timmerman (1981) has been one of the first <strong>to</strong> discuss the <strong>resilience</strong> of society <strong>to</strong> climate<br />

change. In so doing, he links <strong>resilience</strong> <strong>to</strong> <strong>vulnerability</strong>. He defines <strong>resilience</strong> as the measure<br />

of a system’s or part of a system’s capacity <strong>to</strong> absorb <strong>and</strong> recover from the occurrence of a<br />

hazardous event.<br />

Dovers <strong>and</strong> H<strong>and</strong>mer (1992) distinguish between the reactive <strong>and</strong> proactive <strong>resilience</strong> of<br />

society. A society relying on reactive <strong>resilience</strong> approaches the future by strengthening the<br />

status quo <strong>and</strong> making the present system resistant <strong>to</strong> change, whereas one that develops<br />

proactive <strong>resilience</strong> accepts the inevitability of change <strong>and</strong> tries <strong>to</strong> create a system that is<br />

capable of adapting <strong>to</strong> new conditions <strong>and</strong> imperatives. This is an important broadening of<br />

16


the traditional interpretation of <strong>resilience</strong>, which is based on the premise of <strong>resilience</strong> being<br />

tested by an initial perturbation. The distinction made by Dovers <strong>and</strong> H<strong>and</strong>mer (1992) is<br />

based on the major difference between ecosystems <strong>and</strong> societies: the human capacity for anticipation<br />

<strong>and</strong> learning.<br />

Dovers <strong>and</strong> H<strong>and</strong>mer (1992) thus link <strong>resilience</strong> <strong>to</strong> planning <strong>and</strong> <strong>adaptation</strong> <strong>to</strong> hazards. In<br />

a later paper, they develop a typology of institutional <strong>resilience</strong>, which provides a framework<br />

for considering the rigidity <strong>and</strong> inadequacy of present institutional responses <strong>to</strong> global environmental<br />

change (H<strong>and</strong>mer <strong>and</strong> Dovers, 1996). They argue that current institutions <strong>and</strong> policy<br />

processes appear <strong>to</strong> be locked in a type of <strong>resilience</strong> that is characterised by change at<br />

the margins. Responses <strong>to</strong> environmental change are shaped by what is perceived <strong>to</strong> be politically<br />

<strong>and</strong> economically palatable in the near term rather than by the nature <strong>and</strong> scale of the<br />

threat itself.<br />

This type of <strong>resilience</strong>, as well as a type that is characterised by resistance <strong>to</strong> change,<br />

provides some level of stability in society, although there is a potentially large risk that this<br />

apparent stability is not sustainable <strong>and</strong> could lead <strong>to</strong> collapse if society cannot make the social,<br />

economic <strong>and</strong> political changes necessary for survival. The third type of <strong>resilience</strong> described<br />

by H<strong>and</strong>mer <strong>and</strong> Dovers (1996), one that is characterised by openness <strong>and</strong> <strong>adaptation</strong>,<br />

is more likely <strong>to</strong> deal directly with the underlying causes of environmental problems <strong>and</strong> reduces<br />

<strong>vulnerability</strong> by providing a high degree of flexibility. Its key feature is a readiness <strong>to</strong><br />

adopt new basic operating assumptions <strong>and</strong> institutional structures. However, there is also a<br />

potentially large risk involved in moving <strong>to</strong>wards this type of <strong>resilience</strong>. Change deemed as<br />

necessary could turn out <strong>to</strong> be maladaptive, rendering a large cost <strong>to</strong> society.<br />

Adger (2000) investigates the links between social <strong>resilience</strong> <strong>and</strong> ecological <strong>resilience</strong>. He<br />

follows Timmerman (1981) in his definition of social <strong>resilience</strong>: the ability of human communities<br />

<strong>to</strong> withst<strong>and</strong> external shocks or perturbations <strong>to</strong> their infrastructure, such as environmental<br />

variability or social, economic or political upheaval, <strong>and</strong> <strong>to</strong> recover from such perturbations.<br />

Social <strong>resilience</strong> is measured through proxies of institutional change <strong>and</strong> economic<br />

structure, property rights, access <strong>to</strong> resources <strong>and</strong> demographic change (Adger, 1997).<br />

It is argued by many ecologists that <strong>resilience</strong> is the key <strong>to</strong> sustainable ecosystem management<br />

<strong>and</strong> that diversity enhances <strong>resilience</strong>, stability <strong>and</strong> ecosystem functioning (e.g.,<br />

Schulze <strong>and</strong> Mooney, 1993; Peterson et al., 1998; Chapin et al., 2000). Ecological economists<br />

also argue that <strong>resilience</strong> is the key <strong>to</strong> sustainability in the wider sense (e.g., Common, 1995).<br />

However, Adger (2000) observes that whilst <strong>resilience</strong> is certainly related <strong>to</strong> stability, it is not<br />

clear whether this characteristic is always desirable (cf. H<strong>and</strong>mer <strong>and</strong> Dovers, 1996).<br />

This overview of conceptual development of <strong>resilience</strong> shows that what was once a<br />

straightforward concept used only in mechanics is now a complex multi-interpretable concept<br />

with contested definitions <strong>and</strong> even relevance. Nonetheless, the concept of <strong>resilience</strong> is now<br />

used in a great variety of interdisciplinary work concerned with the interactions between<br />

people <strong>and</strong> nature, including <strong>vulnerability</strong> <strong>and</strong> disaster reduction (ISDR, 2002). The most important<br />

development over the past thirty years is the increasing recognition across the disciplines<br />

that human <strong>and</strong> ecological systems are interlinked <strong>and</strong> that their <strong>resilience</strong> relates <strong>to</strong><br />

the functioning <strong>and</strong> interaction of the systems rather than <strong>to</strong> the stability of their components<br />

or the ability <strong>to</strong> maintain or return <strong>to</strong> some equilibrium state.<br />

This recognition has led <strong>to</strong> the establishment of the Resilience Alliance, a network of<br />

scientists with roots mainly in ecology <strong>and</strong> ecological economics, which aims <strong>to</strong> stimulate<br />

academic research on <strong>resilience</strong> <strong>and</strong> inform the global policy process on sustainable development<br />

(Folke et al., 2002). The Resilience Alliance consistently refers <strong>to</strong> social-ecological systems<br />

<strong>and</strong> defines their <strong>resilience</strong> by considering three distinct dimensions (Carpenter et al.,<br />

2001):<br />

The amount of disturbance a system can absorb <strong>and</strong> still remain within the same state or<br />

domain of attraction;<br />

The degree <strong>to</strong> which the system is capable of self-organisation;<br />

The degree <strong>to</strong> which the system can build <strong>and</strong> increase the capacity for learning <strong>and</strong> <strong>adaptation</strong>.<br />

This definition is an amalgamation of the aforementioned definitions of ecological, social<br />

<strong>and</strong> institutional <strong>resilience</strong>. However, <strong>resilience</strong> remains at the conceptual level, creating the<br />

17


danger that there is a “motherhood <strong>and</strong> apple-pie” view on <strong>resilience</strong> that leaves limited<br />

scope for measurement, testing <strong>and</strong> formalisation. The challenge remains <strong>to</strong> transform the<br />

concept of <strong>resilience</strong> in<strong>to</strong> an operational <strong>to</strong>ol for policy <strong>and</strong> management purposes (Klein et<br />

al., 2002).<br />

4.3. Adaptation<br />

Like <strong>vulnerability</strong>, <strong>adaptation</strong> is a term that was given prominence in the UNFCCC as one of<br />

the two response strategies <strong>to</strong> climate change, along with mitigation. Mitigation comprises all<br />

human activities aimed at reducing the emissions or enhancing the sinks of greenhouse gases<br />

such as carbon dioxide, methane <strong>and</strong> nitrous oxide. Adaptation in the context of climate<br />

change refers <strong>to</strong> any adjustment that takes place in natural or human systems in response <strong>to</strong><br />

actual or expected impacts of climate change, aimed at moderating harm or exploiting beneficial<br />

opportunities.<br />

Despite the fact that the UNFCCC refers <strong>to</strong> both mitigation <strong>and</strong> <strong>adaptation</strong>, national <strong>and</strong><br />

international climate policies <strong>to</strong> date have mainly focused on mitigation. In part this reflects<br />

the uncertainty about whether climate change is caused by human activity, which existed until<br />

the publication of the IPCC Second Assessment Report in 1996. It also reflects the lack of<br />

theoretical <strong>and</strong> practical knowledge about <strong>adaptation</strong> <strong>to</strong> climate change, which in turn resulted<br />

from the limited attention given <strong>to</strong> <strong>adaptation</strong> by the scientific community. In his review<br />

of the IPCC Second Assessment Report, Kates (1997) suggests the reason for this limited<br />

attention lies in the existence of two distinct schools of thought about climate change, both<br />

of which have chosen not <strong>to</strong> engage in <strong>adaptation</strong> research.<br />

On the one extreme Kates identifies the “preventionist” school, which argues that the<br />

ongoing increase of atmospheric greenhouse gas concentrations could be catastrophic <strong>and</strong><br />

that drastic action is required <strong>to</strong> reduce emissions. Preventionists fear that increased emphasis<br />

on <strong>adaptation</strong> will weaken society’s willingness <strong>to</strong> reduce emissions <strong>and</strong> thus delay or diminish<br />

mitigation efforts. On the other extreme one finds what Kates refers <strong>to</strong> as the “<strong>adaptation</strong>ist”<br />

school, which sees no need <strong>to</strong> focus on either <strong>adaptation</strong> or mitigation. Adaptationists<br />

argue that natural <strong>and</strong> human systems have a long his<strong>to</strong>ry of adapting naturally <strong>to</strong> changing<br />

circumstances <strong>and</strong> that active <strong>adaptation</strong> would constitute interference with these systems,<br />

bringing with it high social costs.<br />

Following the publication of the IPCC Second Assessment Report, a distinct third school<br />

of thought has emerged, which has been labelled the “realist” school by Klein <strong>and</strong> MacIver<br />

(1999). The realist school positions itself in between the two extreme views of the preventionists<br />

<strong>and</strong> <strong>adaptation</strong>ists. Realists regard climate change as a fact but acknowledge that impacts<br />

are still uncertain. Furthermore, realists appreciate that the planning <strong>and</strong> implementation<br />

of effective <strong>adaptation</strong> options takes time. Therefore, they underst<strong>and</strong> that a process<br />

must be set in motion <strong>to</strong> consider <strong>adaptation</strong> as a crucial <strong>and</strong> realistic response option along<br />

with mitigation (e.g., Parry et al., 1998; Pielke, 1998).<br />

There are various ways <strong>to</strong> classify or distinguish between <strong>adaptation</strong> options. First, depending<br />

on the timing, goal <strong>and</strong> motive of its implementation, <strong>adaptation</strong> can be either reactive<br />

or anticipa<strong>to</strong>ry. Reactive <strong>adaptation</strong> occurs after the initial impacts of climate change<br />

have become manifest, whilst anticipa<strong>to</strong>ry (or proactive) <strong>adaptation</strong> takes place before impacts<br />

are apparent. A second distinction can be based on the system in which the <strong>adaptation</strong><br />

takes place: the natural system (in which <strong>adaptation</strong> is by definition reactive) or the human<br />

system (in which both reactive <strong>and</strong> anticipa<strong>to</strong>ry <strong>adaptation</strong> are observed). Within the human<br />

system, a third distinction can be based on whether the <strong>adaptation</strong> decision is motivated by<br />

private or public interests. Private decision-makers include both individual households <strong>and</strong><br />

commercial companies, whilst public interests are served by governments at all levels. Figure<br />

2 shows examples of <strong>adaptation</strong> activities for each of the five types of <strong>adaptation</strong> that have<br />

thus been defined.<br />

In addition <strong>to</strong> the ones made above, other <strong>adaptation</strong> distinctions are discussed in a paper<br />

by Smit et al. (2000), which is presented later in this thesis. A useful distinction that is<br />

often made is the one between planned <strong>and</strong> au<strong>to</strong>nomous <strong>adaptation</strong> (Carter et al., 1994).<br />

Planned <strong>adaptation</strong> is the result of a deliberate policy decision that is based on an awareness<br />

that conditions have changed or are about <strong>to</strong> change <strong>and</strong> that action is required <strong>to</strong> return <strong>to</strong>,<br />

maintain or achieve a desired state. Au<strong>to</strong>nomous <strong>adaptation</strong> involves the changes that natural<br />

<strong>and</strong> most human systems will undergo in response <strong>to</strong> changing conditions, irrespective of any<br />

18


policy plan or decision. Instead, au<strong>to</strong>nomous <strong>adaptation</strong> will be triggered by market or welfare<br />

changes induced by climate change. Au<strong>to</strong>nomous <strong>adaptation</strong> in human systems would<br />

therefore be in the ac<strong>to</strong>r’s rational self-interest, whilst the focus of planned <strong>adaptation</strong> is on<br />

collective needs (Leary, 1999). Thus defined, au<strong>to</strong>nomous <strong>and</strong> planned <strong>adaptation</strong> largely correspond<br />

with private <strong>and</strong> public <strong>adaptation</strong>, respectively (see Figure 2).<br />

Anticipa<strong>to</strong>ry<br />

Reactive<br />

Natural<br />

Systems<br />

· Changes in length of growing season<br />

· Changes in ecosystem composition<br />

·Wetl<strong>and</strong> migration<br />

Human<br />

Systems<br />

Private<br />

Public<br />

· Purchase of insurance<br />

· Construction of houses on stilts<br />

·Redesign of oil-rigs<br />

· Early-warning systems<br />

· New building codes, design st<strong>and</strong>ards<br />

· Incentives for relocation<br />

·Changes in farm practices<br />

· Changes in insurance premiums<br />

· Purchase of air-conditioning<br />

·Compensa<strong>to</strong>ry payments, subsidies<br />

· Enforcement of building codes<br />

· Beach nourishment<br />

Figure 2 — Matrix showing the five prevalent types of <strong>adaptation</strong> <strong>to</strong> climate change, including<br />

examples (based on Klein, 1998).<br />

Article 3.3 of the UNFCCC suggests that anticipa<strong>to</strong>ry planned <strong>adaptation</strong> (as well as<br />

mitigation) deserves particular attention from the international climate change community:<br />

“The Parties should take precautionary measures <strong>to</strong> anticipate, prevent or minimise the causes of climate<br />

change <strong>and</strong> mitigate its adverse effects. Where there are threats of serious or irreversible damage, lack<br />

of full scientific certainty should not be used as a reason for postponing such measures, taking in<strong>to</strong> account<br />

that policies <strong>and</strong> measures <strong>to</strong> deal with climate change should be cost-effective so as <strong>to</strong> ensure<br />

global benefits at the lowest possible cost. (...).”<br />

Anticipa<strong>to</strong>ry <strong>adaptation</strong> is aimed at reducing a system’s <strong>vulnerability</strong> by either minimising<br />

risk or maximising adaptive capacity. Five generic objectives of anticipa<strong>to</strong>ry <strong>adaptation</strong><br />

can be identified (Klein <strong>and</strong> Tol, 1997; Klein, 2001):<br />

Increasing robustness of infrastructural designs <strong>and</strong> long-term investments—for example<br />

by extending the range of temperature or precipitation a system can withst<strong>and</strong> without<br />

failure <strong>and</strong>/or changing a system’s <strong>to</strong>lerance of loss or failure (e.g., by increasing economic<br />

reserves or insurance);<br />

Increasing flexibility of vulnerable managed systems—for example by allowing mid-term<br />

adjustments (including change of activities or location) <strong>and</strong>/or reducing economic lifetimes<br />

(including increasing depreciation);<br />

Enhancing adaptability of vulnerable natural systems—for example by reducing other<br />

(non-climatic) stresses <strong>and</strong>/or removing barriers <strong>to</strong> migration (such as establishing ecocorridors);<br />

Reversing trends that increase <strong>vulnerability</strong> (“mal<strong>adaptation</strong>”)—for example by introducing<br />

zoning regulation in vulnerable areas such as floodplains <strong>and</strong> coastal zones;<br />

Improving societal awareness <strong>and</strong> preparedness—for example by informing the public of<br />

the risks <strong>and</strong> possible consequences of climate change <strong>and</strong>/or setting up early-warning<br />

systems.<br />

Each of these five objectives of <strong>adaptation</strong> is relevant for coastal zones. However, for<br />

coastal zones another classification of <strong>adaptation</strong> options is often used. This classification, introduced<br />

by IPCC CZMS (1990) <strong>and</strong> still the basis of many coastal <strong>adaptation</strong> analyses, distinguishes<br />

between the following three basic strategies:<br />

19


Protect—<strong>to</strong> reduce the risk of the event by decreasing the probability of its occurrence;<br />

Accommodate—<strong>to</strong> increase society’s ability <strong>to</strong> cope with the effects of the event;<br />

Retreat—<strong>to</strong> reduce the risk of the event by limiting its potential effects.<br />

For each of these strategies Bijlsma et al. (1996) provides a list of different options,<br />

ranging from hard structural options <strong>and</strong> strict regulation of hazard zones <strong>to</strong> emergency planning<br />

<strong>and</strong> managed realignment. Klein <strong>and</strong> Nicholls (1998) then analyse the process <strong>and</strong> timing<br />

of their implementation, whilst Klein et al. (2000, 2001) discuss the three strategies in detail<br />

<strong>and</strong> provide examples of technologies for implementing each of them.<br />

The three coastal <strong>adaptation</strong> strategies roughly coincide with the first three of the five<br />

<strong>adaptation</strong> objectives listed above. Protecting coastal zones against sea-level rise <strong>and</strong> other<br />

climatic changes would involve increasing the robustness of infrastructural designs <strong>and</strong> longterm<br />

investments such as seawalls <strong>and</strong> other coastal infrastructure. A strategy <strong>to</strong> accommodate<br />

sea-level rise could include increasing the flexibility of managed systems such as agriculture,<br />

<strong>to</strong>urism <strong>and</strong> human settlements in coastal zones. A retreat strategy would serve <strong>to</strong> enhance<br />

the adaptability (or <strong>resilience</strong>) of coastal wetl<strong>and</strong>s, by allowing them space <strong>to</strong> migrate<br />

<strong>to</strong> higher l<strong>and</strong> as sea level rises.<br />

Policies <strong>and</strong> practices that are unrelated <strong>to</strong> climate but which do increase a system’s <strong>vulnerability</strong><br />

<strong>to</strong> climate change are termed mal<strong>adaptation</strong> (Bur<strong>to</strong>n, 1996, 1997). Examples of<br />

mal<strong>adaptation</strong> in coastal zones include investments in hazardous zones, inappropriate<br />

coastal-defence schemes, s<strong>and</strong> or coral mining <strong>and</strong> coastal-habitat conversions. A common<br />

cause of mal<strong>adaptation</strong> is a lack of information on the potential external effects of proposed<br />

developments on other sec<strong>to</strong>rs, or a lack of consideration thereof. More proactive <strong>and</strong> integrated<br />

planning <strong>and</strong> management of coastal zones is widely suggested as an effective mechanism<br />

for strengthening sustainable development (e.g., Cicin-Sain, 1993; Ehler et al., 1997;<br />

Cicin-Sain <strong>and</strong> Knecht, 1998; see Section 5).<br />

Empirical information on coastal <strong>adaptation</strong> <strong>to</strong> climate change is still scarce, so uncertainty<br />

remains considerable. Continued impact <strong>and</strong> <strong>adaptation</strong> assessment, combined with<br />

fundamental research on coastal system response <strong>and</strong> economic, institutional, legal <strong>and</strong><br />

socio-cultural aspects of <strong>adaptation</strong>, are required <strong>to</strong> underst<strong>and</strong> which <strong>adaptation</strong> options<br />

might be most appropriate <strong>and</strong> most effectively implemented. The synthesis chapter of this<br />

thesis will propose ways <strong>to</strong> improve <strong>adaptation</strong> assessment in coastal zones, based on the<br />

concept of adaptive capacity.<br />

5. <strong>Coastal</strong> <strong>vulnerability</strong>, <strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong> <strong>to</strong> climate change: the<br />

policy <strong>and</strong> management context<br />

Although the primary goal of the aforementioned <strong>vulnerability</strong> assessments has been <strong>to</strong> identify<br />

a country’s coastal <strong>vulnerability</strong> <strong>to</strong> climate change <strong>and</strong> associated sea-level rise, many of<br />

the assessments have shown that climate change is often not the most crucial issue for a<br />

coastal region. This is particularly true in areas where pressures from population growth <strong>and</strong><br />

economic development are already creating problems <strong>and</strong> hazards. Nonetheless, as stated before,<br />

these current pressures may have adversely affected the coastal ecosystem’s <strong>resilience</strong><br />

<strong>and</strong> thereby its ability <strong>to</strong> cope with additional pressures such as climate change <strong>and</strong> sea-level<br />

rise. For example, the conversion of coastal wetl<strong>and</strong>s <strong>to</strong> agricultural l<strong>and</strong> can be at the expense<br />

of the coast’s ability <strong>to</strong> provide a buffering capacity <strong>and</strong> thus protect the l<strong>and</strong> against<br />

the dynamics of the sea. As climate changes <strong>and</strong> sea level rises, this capacity will become<br />

even more important than it is <strong>to</strong>day, preventing coastal areas from being eroded <strong>and</strong> inundated.<br />

This illustrates that when current coastal pressures are not adequately dealt with in<br />

the short term, coastal zones will become increasingly more vulnerable <strong>to</strong> the consequences<br />

of climate change <strong>and</strong> sea-level rise (WCC’93, 1994).<br />

In spite of many local <strong>and</strong> national efforts, traditional approaches <strong>to</strong> the management<br />

<strong>and</strong> use of coastal resources have often proved <strong>to</strong> be insufficient <strong>to</strong> achieve sustainable development<br />

(Cicin-Sain <strong>and</strong> Knecht, 1998). As a result, coastal resources are being degraded<br />

<strong>and</strong> lost in many parts of the world. Some policy process is needed that can address the resource-use<br />

conflicts discussed in Section 2, as well as find the balance between short-term<br />

economic <strong>and</strong> longer-term environmental interests. In other words, a process is needed <strong>to</strong><br />

20


implement decisions on the mix of uses that best serves the needs of society now <strong>and</strong> in the<br />

future.<br />

Integrated coastal zone management (ICZM) has been widely recognised <strong>and</strong> promoted as<br />

the most appropriate process <strong>to</strong> deal with these current <strong>and</strong> long-term coastal challenges. According<br />

<strong>to</strong> WCC’93 (1994), Ehler et al. (1997) <strong>and</strong> Cicin-Sain <strong>and</strong> Knecht (1998), ICZM provides<br />

coastal societies with an opportunity <strong>to</strong> move <strong>to</strong>wards sustainable development. Integrated<br />

management of conflicting uses <strong>and</strong> activities is essential for this goal. Moreover, ICZM allows<br />

current <strong>and</strong> future interests in coastal areas <strong>and</strong> resources <strong>to</strong> be taken in<strong>to</strong> account. By considering<br />

short, medium <strong>and</strong> long-term interests, ICZM stimulates economic development of<br />

coastal areas <strong>and</strong> resources, whilst reducing the degradation of their natural systems. Third,<br />

ICZM provides a framework within which flexible responses can be developed <strong>to</strong> deal with the<br />

inherent uncertainty about the future, including rates <strong>and</strong> magnitude of climate change. In<br />

short, ICZM can provide coastal states with a process <strong>to</strong> enhance economic development <strong>and</strong><br />

improve the quality of life (WCC’93, 1994).<br />

At the World Coast Conference, ICZM was defined as follows:<br />

Integrated coastal zone management involves the comprehensive assessment, setting of objectives, planning<br />

<strong>and</strong> management of coastal systems <strong>and</strong> resources, taking in<strong>to</strong> account traditional, cultural <strong>and</strong> his<strong>to</strong>rical<br />

perspectives <strong>and</strong> conflicting interests <strong>and</strong> uses; it is a continuous <strong>and</strong> evolutionary process for<br />

achieving sustainable development.<br />

The purpose of ICZM is <strong>to</strong> provide the conditions that will facilitate development, stimulate<br />

progress <strong>and</strong> encourage (changes in) human <strong>and</strong> institutional behaviour in order <strong>to</strong><br />

achieve shared goals. In general, these goals are specified targets related <strong>to</strong> the desired mix<br />

of goods, services <strong>and</strong> values <strong>to</strong> be produced, consumed or conserved. ICZM must anticipate<br />

<strong>and</strong> respond <strong>to</strong> the needs of coastal communities; public participation <strong>and</strong> stakeholder involvement<br />

in the planning <strong>and</strong> implementation of ICZM are therefore essential.<br />

To be successful, ICZM should include the following (Bijlsma et al., 1996):<br />

Integration of programmes <strong>and</strong> plans for economic development, environmental management<br />

<strong>and</strong> l<strong>and</strong> use;<br />

Integration of programmes for economic sec<strong>to</strong>rs, such as agriculture, fisheries, energy,<br />

transportation, water resources, waste disposal <strong>and</strong> <strong>to</strong>urism;<br />

Integration of responsibilities for various tasks of management amongst levels of government<br />

(local, state/provincial, national <strong>and</strong> regional), as well as between the public <strong>and</strong><br />

private sec<strong>to</strong>r;<br />

Integration of available resources for management (personnel, funds, materials <strong>and</strong><br />

equipment);<br />

Integration amongst scientific disciplines, such as ecology, geomorphology, hydrology,<br />

economics, engineering, political science, behavioural science <strong>and</strong> law.<br />

Although integration in management requires more extensive analysis, preparation <strong>and</strong><br />

planning than is common in sec<strong>to</strong>ral management, the overall costs of ICZM can ultimately be<br />

lower than the cumulative costs of separate sec<strong>to</strong>ral approaches. In addition, taking a proactive,<br />

or precautionary, approach <strong>to</strong> ICZM (i.e., acting before irreversible damage is realised)<br />

can provide not only environmental but also economic benefits (Vellinga <strong>and</strong> Leatherman,<br />

1989; Tol et al., 1995).<br />

This chapter has shown that <strong>adaptation</strong> <strong>to</strong> climate change cannot be separated from the<br />

planning <strong>and</strong> management efforts <strong>to</strong> address present-day problems in coastal areas. Carefully<br />

harmonising the different sec<strong>to</strong>ral development options <strong>and</strong> needs is a first step in an evolutionary<br />

process <strong>to</strong>wards sustainable development in coastal zones. Sustainable development<br />

has, by definition, an extended time horizon. Long-term thinking, as encouraged by the consideration<br />

of climate change, is therefore a key component of ICZM <strong>and</strong> can be economically<br />

feasible (Tol et al., 1995).<br />

However, it is important <strong>to</strong> note that there is no single recipe for successful ICZM. Every<br />

coastal zone faces different challenges, which require unique approaches <strong>to</strong> management.<br />

Whereas one can identify basic steps in ICZM (problem recognition, analysis <strong>and</strong> planning, implementation<br />

of measures, operation <strong>and</strong> maintenance <strong>and</strong> moni<strong>to</strong>ring <strong>and</strong> evaluation of the<br />

effectiveness of the measures), the most appropriate management approach will depend <strong>to</strong> a<br />

21


large extent on cultural, political, economic <strong>and</strong> his<strong>to</strong>rical conditions, <strong>and</strong> will therefore vary<br />

between <strong>and</strong> within countries.<br />

An important difference between sec<strong>to</strong>ral <strong>and</strong> integrated coastal zone management is<br />

the interaction <strong>and</strong> collaboration between different stakeholders so as <strong>to</strong> ensure the careful<br />

consideration of both public <strong>and</strong> private interests. The next two sections provide a brief overview<br />

of public <strong>and</strong> private sec<strong>to</strong>r responsibilities in ICZM, with a particular focus on <strong>adaptation</strong><br />

<strong>to</strong> climate change.<br />

5.1. Public initiatives <strong>and</strong> responsibilities<br />

The public sec<strong>to</strong>r represents governments from the local <strong>to</strong> the national level, as well as international<br />

institutional organisations. The public sec<strong>to</strong>r is typically concerned with meeting<br />

collective needs <strong>and</strong> safeguarding collective interests (such as health, safety <strong>and</strong> biodiversity),<br />

as well as facilitating private sec<strong>to</strong>r activities within a regula<strong>to</strong>ry, fiscal <strong>and</strong> planning<br />

framework designed <strong>to</strong> prevent or minimise social costs. In addition, the public sec<strong>to</strong>r can initiate<br />

activities aimed at awareness raising, capacity building <strong>and</strong> technology transfer (Klein et<br />

al., 2001).<br />

Many ecosystem goods <strong>and</strong> services listed in Table 1 are so-called commons: they are<br />

available at no cost <strong>to</strong> anyone who wants <strong>to</strong> use them. The fact that they do not have a market<br />

price does not mean they do not have a value (see Section 2.3.3). However, this value is<br />

often only appreciated once the ecosystem good or service has been overexploited <strong>and</strong> lost.<br />

The increasing concentration of people <strong>and</strong> activities in coastal zones has increased the potential<br />

for multiple-use conflicts, as the activities of one stakeholder impinge on the interests<br />

or activities of another stakeholder.<br />

As stated in Section 2.1, coastal zone management is traditionally considered as a process<br />

<strong>to</strong> h<strong>and</strong>le conflicts between the various (potential) users of increasingly scarce coastal resources<br />

<strong>and</strong> <strong>to</strong> address current problems that result from stakeholders pursuing their own sec<strong>to</strong>ral<br />

interests. As such, coastal zone management has focused strongly on conflicts stemming<br />

from the multiple uses of resources provided by user <strong>and</strong> production functions. The United<br />

States <strong>Coastal</strong> Zone Management Act of 1972 was one of the first public sec<strong>to</strong>r initiatives<br />

worldwide <strong>to</strong> co-ordinate sec<strong>to</strong>ral activities in coastal zones. An important component of the<br />

act is an agreement between federal <strong>and</strong> state governments, whereby the federal government<br />

offers financial <strong>and</strong> technical assistance <strong>to</strong> states. States can choose <strong>to</strong> develop <strong>and</strong> implement<br />

coastal zone management programmes, but they must adhere <strong>to</strong> minimum federal<br />

st<strong>and</strong>ards on protecting coastal resources, ensuring public access <strong>to</strong> the coast, managing development<br />

along the coast <strong>and</strong> managing coastal hazards (Kay <strong>and</strong> Alder, 1999).<br />

As pressures on coasts continued <strong>to</strong> increase, the need for ICZM was recognised more<br />

widely, not only at a national level but internationally as well. In the 1980s <strong>and</strong> 1990s, a<br />

range of international organisations promoted the development of ICZM, including UNEP, the<br />

World Bank, the Organisation for Economic Co-operation <strong>and</strong> Development (OECD) <strong>and</strong> the International<br />

Union for the Conservation of Nature (IUCN). At the United Nations Conference on<br />

Environment <strong>and</strong> Development (UNCED) in 1992, the management of coastal zones featured<br />

prominently in the conference’s main product, Agenda 21. Chapter 17 of Agenda 21 stresses<br />

both the importance of oceans <strong>and</strong> coasts for life on earth <strong>and</strong> the positive opportunity for<br />

sustainable development represented by oceans <strong>and</strong> coastal zones. It states that new approaches<br />

<strong>to</strong> marine <strong>and</strong> coastal management are needed; approaches that are integrated in<br />

content <strong>and</strong> precautionary <strong>and</strong> anticipa<strong>to</strong>ry in ambit. Implementation of the recommendations<br />

in Chapter 17 involves a multitude of ac<strong>to</strong>rs, ranging from local grassroots organisations<br />

<strong>to</strong> global institutions <strong>and</strong> everything in between (Cicin-Sain et al., 1995).<br />

Another major product of UNCED was the United Nations Framework Convention on Climate<br />

Change (see Section 4.1). In Article 4.1(e), Parties <strong>to</strong> the UNFCCC commit themselves <strong>to</strong><br />

develop <strong>and</strong> elaborate appropriate <strong>and</strong> integrated plans for coastal zone management in a<br />

process of preparing for <strong>adaptation</strong> <strong>to</strong> the impacts of climate change. This commitment has<br />

added a new dimension <strong>to</strong> ICZM. Whereas traditional ICZM focused on coastal development<br />

pressures, multiple-use conflicts <strong>and</strong> resource degradation, ICZM now includes a long-term<br />

sustainability component.<br />

The World Coast Conference, held in November 1993 in Noordwijk, The Netherl<strong>and</strong>s, was<br />

held in response <strong>to</strong> Chapter 17 of Agenda 21 <strong>and</strong> Article 4.1(e) of the UNFCCC. The conference<br />

brought <strong>to</strong>gether experts, decision-makers <strong>and</strong> representatives of non-governmental or-<br />

22


ganisations <strong>to</strong> consider the relevance of coastal <strong>vulnerability</strong> <strong>to</strong> climate change <strong>to</strong> coastal<br />

zone management. Two of its objectives were <strong>to</strong> provide an opportunity for coastal countries<br />

<strong>to</strong> exchange information <strong>and</strong> experiences in assessing <strong>vulnerability</strong> <strong>to</strong> climate change <strong>and</strong> in<br />

developing coastal zone management plans <strong>and</strong> <strong>to</strong> contribute <strong>to</strong> the development of common<br />

concepts, techniques <strong>and</strong> <strong>to</strong>ols for these two activities. The conference succeeded in building<br />

awareness of the importance of ICZM in reducing coastal <strong>vulnerability</strong> <strong>to</strong> climate change <strong>and</strong><br />

the opportunities it presents (e.g., Bijlsma et al., 1996). One remaining challenge, however,<br />

is <strong>to</strong> translate the words of global initiatives in<strong>to</strong> local action (see also the synthesis chapter<br />

of this thesis).<br />

5.2. Private initiatives <strong>and</strong> responsibilities<br />

Private-sec<strong>to</strong>r interests in coastal zones are very diverse <strong>and</strong> include agriculture, fisheries,<br />

transportation, mining, manufacturing <strong>and</strong> <strong>to</strong>urism. In addition, space is required for human<br />

habitation <strong>and</strong> nature conservation. Private decision-makers include both individual households<br />

<strong>and</strong> commercial companies. Their actions are triggered by market or welfare changes<br />

<strong>and</strong> are aimed at maximising their profit or utility (an economic concept referring <strong>to</strong> the degree<br />

of personal satisfaction an individual derives from consuming some quantity of a good or<br />

service at a particular point in time). Different companies <strong>and</strong> individuals have different objectives<br />

<strong>and</strong> attitudes <strong>to</strong> the use of coastal resources, which may lead <strong>to</strong> conflicts.<br />

As stated before, ICZM is primarily a mechanism <strong>to</strong> avoid such multiple-use conflicts between<br />

sec<strong>to</strong>ral stakeholders. Whilst the responsibility for initiating an ICZM programme lies<br />

with the public sec<strong>to</strong>r, success depends largely on the will <strong>and</strong> motivation of the private sec<strong>to</strong>r<br />

<strong>to</strong> adopt the objectives <strong>and</strong> principles of the programme. Therefore, from the moment<br />

the initiative <strong>to</strong>wards developing an ICZM programme is taken, a continuous dialogue with<br />

private sec<strong>to</strong>r stakeholders is essential as part of an iterative process aimed at incorporating<br />

stakeholders’ concerns <strong>and</strong> developing broad underst<strong>and</strong>ing <strong>and</strong> support for the needs, means<br />

<strong>and</strong> ends of the programme.<br />

The private sec<strong>to</strong>r tends <strong>to</strong> have a relatively short planning horizon, which is not conducive<br />

<strong>to</strong> considering potential effects of climate change on its activities. Therefore, the public<br />

sec<strong>to</strong>r has the strongest <strong>and</strong> most direct incentives <strong>to</strong> adapt <strong>to</strong> climate change, although sealevel<br />

rise <strong>and</strong> its impacts can be a direct threat <strong>to</strong> the profitability of particularly the <strong>to</strong>urist<br />

sec<strong>to</strong>r <strong>and</strong> ports <strong>and</strong> harbours. Nonetheless, the private sec<strong>to</strong>r typically does not take the initiative<br />

for coastal <strong>adaptation</strong> <strong>to</strong> climate change, because benefits are small or uncertain, <strong>and</strong><br />

action is expected from the government <strong>to</strong> protect private sec<strong>to</strong>r interests (Klein et al.,<br />

2001). In developing countries, the private sec<strong>to</strong>r is generally a less significant economic<br />

force; so again, governments are expected <strong>to</strong> lead the way in coastal <strong>adaptation</strong>. However, in<br />

government-initiated coastal <strong>adaptation</strong>, the private sec<strong>to</strong>r is often involved in the planning,<br />

design <strong>and</strong> implementation of <strong>adaptation</strong> measures.<br />

Potentially successful public-private partnerships for ICZM are being developed particularly<br />

at a local level. For example, in the KERN region in Schleswig-Holstein in Germany (Kiel,<br />

Eckernförde, Rendsburg <strong>and</strong> Neumünster), local <strong>and</strong> district government, the chamber of<br />

commerce <strong>and</strong> academic institutes are working <strong>to</strong>gether <strong>to</strong> implement an ICZM programme<br />

that aims <strong>to</strong> strike a balance between economic <strong>and</strong> ecological interests. The programme<br />

does not yet have statu<strong>to</strong>ry powers, but the progress made so far is evidence that the process<br />

of consensus building does not require a formal institutional framework. A challenge for the<br />

next few years will be <strong>to</strong> translate the consensus in<strong>to</strong> action.<br />

6. Scope, methodology <strong>and</strong> structure of the thesis<br />

As explained in the introduction <strong>to</strong> this chapter, the work presented in this PhD thesis has not<br />

been carried out within a single, well-defined project. It combines <strong>and</strong> integrates results of a<br />

number of different studies commissioned <strong>and</strong> funded by a range of organisations. Using a<br />

consistent <strong>and</strong> academically sound analytical framework <strong>to</strong> design these different studies <strong>and</strong><br />

present their results was usually not the clients’ most important objective. Nonetheless, the<br />

thesis presents a consistent approach <strong>to</strong> increasing the underst<strong>and</strong>ing of the three key concepts<br />

of <strong>vulnerability</strong>, <strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong> <strong>to</strong> climate change. This section introduces<br />

23


the central question underpinning the work presented in this thesis <strong>and</strong> discusses the methodology<br />

pursued <strong>to</strong>wards answering this question. It then introduces the six papers <strong>and</strong> the<br />

synthesis chapter that constitute the remainder of the thesis. Finally, this section indicates<br />

the parts of the multi-authored papers for which the PhD c<strong>and</strong>idate is responsible.<br />

6.1. Problem definition <strong>and</strong> methodology<br />

The earlier parts of this chapter showed that coastal zones provide many goods <strong>and</strong> services<br />

crucial <strong>to</strong> human well-being. The economic opportunities offered by these goods <strong>and</strong> services<br />

have attracted many people <strong>and</strong> investments <strong>to</strong> the coast, leading <strong>to</strong> the unsustainable use<br />

<strong>and</strong> unrestricted development of coastal resources. Many coastal areas have been so heavily<br />

modified <strong>and</strong> intensively developed that their natural <strong>resilience</strong> <strong>to</strong> the effects of climate<br />

change has been significantly reduced.<br />

The importance of coastal resources <strong>to</strong> human society <strong>and</strong> the risk <strong>to</strong> which they may be<br />

exposed from climate change have triggered studies <strong>to</strong> assess the <strong>vulnerability</strong> of coastal<br />

zones <strong>to</strong> climate change. One of the aims of these studies has been <strong>to</strong> advise coastal planners<br />

<strong>and</strong> managers on the need <strong>and</strong> opportunity <strong>to</strong> implement <strong>adaptation</strong> options <strong>to</strong> reduce potential<br />

impacts. Results of these studies were presented in Section 3.2.<br />

It can be argued that whilst these studies have raised awareness of the potential impacts<br />

of climate change on coastal zones <strong>and</strong> thus contributed <strong>to</strong> the international climate policy<br />

process, their usefulness for <strong>adaptation</strong> has been limited, especially in developing countries.<br />

Reasons for this limited usefulness include:<br />

Limited availability of relevant natural <strong>and</strong> socio-economic data;<br />

A bias of study teams <strong>to</strong>wards the natural sciences;<br />

A lack of methodological guidance on <strong>adaptation</strong> assessment;<br />

A mismatch between the scale of the <strong>vulnerability</strong> assessment <strong>and</strong> the scale at which <strong>adaptation</strong><br />

options are implemented;<br />

A mismatch between the institutions involved in <strong>vulnerability</strong> assessment <strong>and</strong> those responsible<br />

for facilitating <strong>and</strong> implementing <strong>adaptation</strong> options.<br />

At the root of these five reasons lies a general lack of underst<strong>and</strong>ing of what <strong>adaptation</strong> is<br />

<strong>and</strong> how it could help <strong>to</strong> reduce <strong>vulnerability</strong> <strong>to</strong> climate change. Until recently, opportunities<br />

<strong>to</strong> learn from experience gained in natural hazard reduction <strong>and</strong> natural resource management,<br />

which have a tradition of adapting <strong>to</strong> climate variability, were not recognised.<br />

The work that is the basis for this PhD thesis has aimed at contributing <strong>to</strong> improved<br />

assessments of coastal <strong>vulnerability</strong> <strong>to</strong> climate change by developing a stronger conceptual<br />

<strong>and</strong> methodological basis for underst<strong>and</strong>ing the process of <strong>adaptation</strong>. It is a first step in an<br />

effort <strong>to</strong> ensure greater policy relevance of <strong>vulnerability</strong> assessments, as well as <strong>to</strong> improve<br />

the academic rigour <strong>and</strong> validity of assessments.<br />

Most of the work presented in this thesis is of a conceptual nature, using literature <strong>and</strong><br />

case studies <strong>to</strong> develop new insights <strong>and</strong> ideas. Rather than applying a particular existing<br />

methodology based on data collection <strong>and</strong> empirical analysis, this PhD thesis provides the<br />

building blocks of a new methodology, which is yet <strong>to</strong> be applied <strong>and</strong> tested (see also the final<br />

chapter of this thesis).<br />

6.2. Structure of the thesis<br />

In hindsight, the coastal chapter of the UNEP H<strong>and</strong>book on Methods for Climate Change Impact<br />

Assessment <strong>and</strong> Adaptation Strategies (Klein <strong>and</strong> Nicholls, 1998) was the starting point<br />

for the conceptual <strong>and</strong> analytical thinking on coastal <strong>vulnerability</strong> <strong>to</strong> climate change. The paper<br />

by Klein <strong>and</strong> Nicholls (1999) provides the theoretical underpinning of this UNEP chapter,<br />

along with a framework for <strong>vulnerability</strong> assessment. This paper is the first of six peer-reviewed<br />

academic papers that form the core of this thesis <strong>and</strong> can be seen as setting out the<br />

conceptual challenges explored in subsequent work.<br />

The second paper in this thesis (Klein et al., 1998) focuses on coastal <strong>resilience</strong>. It was a<br />

by-product of a study aimed at defining <strong>and</strong> making operational the concept of <strong>resilience</strong> for<br />

Dutch coastal management in the light of climate change <strong>and</strong> sea-level rise. This study was<br />

24


conducted with Delft Hydraulics for the Dutch Ministry of Transport, Public Works <strong>and</strong> Water<br />

Management. In distinguishing between morphological, ecological <strong>and</strong> socio-economic <strong>resilience</strong>,<br />

the paper shows the importance of taking an interdisciplinary approach <strong>to</strong> assessing<br />

coastal <strong>vulnerability</strong>, <strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong> <strong>to</strong> climate change.<br />

The third paper (Klein <strong>and</strong> Bateman, 1998) presents the only empirical work in this thesis.<br />

It is an environmental economic analysis of the value of recreation in a coastal freshwater<br />

nature reserve in North Norfolk, United Kingdom. The Norfolk Wildlife Trust used the results<br />

<strong>to</strong> argue against a recommendation in the Shoreline Management Plan for North Norfolk,<br />

which was <strong>to</strong> adapt <strong>to</strong> sea-level rise by implementing a policy of managed retreat. Such a<br />

policy would most likely lead <strong>to</strong> the disappearance of the nature reserve. The Norfolk Wildlife<br />

Trust encouraged the inclusion of the results of the environmental economic analysis in the<br />

cost-benefit analysis on which the recommendation for managed retreat was based.<br />

The fourth paper (Smit et al., 2000) is the only paper in this thesis that does not have a<br />

coastal focus. This paper was prepared <strong>to</strong> provide structure <strong>to</strong> the then emerging academic<br />

<strong>and</strong> policy debate on the desirability, feasibility <strong>and</strong> “assessability” of <strong>adaptation</strong> <strong>to</strong> climate<br />

change. Many of the concepts introduced in this paper, including the need <strong>to</strong> consider natural<br />

climate variability as well as human-induced climate change when defining needs <strong>and</strong> opportunities<br />

for <strong>adaptation</strong>, have been adopted in the IPCC Third Assessment Report.<br />

The process of <strong>adaptation</strong> <strong>and</strong> an approach <strong>to</strong> its assessment are analysed more closely in<br />

the fifth paper in this thesis (Klein et al., 1999). This paper presents coastal management experiences<br />

from The Netherl<strong>and</strong>s, the United Kingdom <strong>and</strong> Japan, which show that <strong>adaptation</strong><br />

<strong>to</strong> natural variability <strong>and</strong> anticipated change is an iterative, cyclical process. Elements in this<br />

cycle are information collection <strong>and</strong> awareness raising, planning <strong>and</strong> design, implementation,<br />

<strong>and</strong> moni<strong>to</strong>ring <strong>and</strong> evaluation. The paper argues that the failure <strong>to</strong> consider all four steps in<br />

<strong>adaptation</strong> assessments has constrained the development of <strong>adaptation</strong> strategies that are<br />

supported by the relevant ac<strong>to</strong>rs <strong>and</strong> integrated in<strong>to</strong> existing management. It introduces the<br />

term “adaptive capacity”, which has become a central focus for linking <strong>adaptation</strong> with development<br />

priorities.<br />

The final paper (Klein et al., 2001) uses the framework developed by Klein et al. (1999)<br />

<strong>to</strong> provide a detailed overview of the types of technologies that are available <strong>to</strong> plan, facilitate<br />

<strong>and</strong> implement <strong>adaptation</strong> in coastal zones. It is based on the coastal <strong>adaptation</strong> chapter<br />

of the IPCC Special Report on Methodological <strong>and</strong> Technological Issues in Technology Transfer<br />

(Klein et al., 2000). This paper again links climate change with climate variability <strong>and</strong> emphasises<br />

the importance of building capacity for technology transfer in developing countries.<br />

The concluding chapter of this thesis presents a synthesis of the six papers in the light of<br />

recent initiatives that build on the conceptual work on <strong>vulnerability</strong>, <strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong>.<br />

It discusses in more detail than the preceding papers the need for <strong>adaptation</strong> in developing<br />

countries <strong>and</strong> the subsequent importance of integrating <strong>adaptation</strong> policy with development<br />

activities. The chapter outlines a number of possible next steps than can be taken <strong>to</strong><br />

advance the academic underst<strong>and</strong>ing of <strong>vulnerability</strong>, <strong>resilience</strong> <strong>and</strong> <strong>adaptation</strong> whilst increasing<br />

the adaptive capacity of developing countries.<br />

6.3. Contribution <strong>to</strong> the papers<br />

The times when great minds could work in isolation on challenging research questions <strong>and</strong><br />

produce meaningful results are long gone. No academic work of quality is possible without extensive<br />

collaboration with other scientists. This is particularly the case for interdisciplinary<br />

research, which explains why all the following six papers are multi-authored. To enable the<br />

Christian-Albrechts-Universität zu Kiel <strong>to</strong> verify whether this PhD thesis contains sufficient<br />

material of quality by the PhD c<strong>and</strong>idate himself, Table 5 gives an overview of his contributions<br />

<strong>to</strong> each of the papers.<br />

25


Paper<br />

Klein <strong>and</strong> Nicholls (1999)<br />

Klein et al. (1998)<br />

Klein <strong>and</strong> Bateman (1998)<br />

Smit et al. (2000)<br />

Klein et al. (1999)<br />

Klein et al. (2001)<br />

Contributions by R.J.T. Klein<br />

This paper was written by Klein, based on the UNEP chapter by Klein <strong>and</strong> Nicholls<br />

(1998). For the UNEP chapter Klein developed the conceptual idea <strong>and</strong> did most of the<br />

writing. Nicholls contributed sections on coastal morphodynamic processes, wetl<strong>and</strong>s<br />

<strong>and</strong> aerial videography-assisted <strong>vulnerability</strong> assessment.<br />

This paper was written by Klein, based partly on contributions from Smit, Goosen <strong>and</strong><br />

Hulsbergen during a workshop.<br />

This award-winning paper was written by Klein, based on his own survey <strong>and</strong> statistical<br />

analyses. Bateman helped <strong>to</strong> develop the survey methodology <strong>and</strong> interpret the results<br />

of the statistical analyses.<br />

This paper was written mainly by Smit, based partly on text inputs from Bur<strong>to</strong>n <strong>and</strong><br />

Klein <strong>and</strong> a literature survey by W<strong>and</strong>el. Klein contributed <strong>to</strong> various parts of the paper,<br />

particularly <strong>to</strong> the typology of <strong>adaptation</strong>.<br />

This paper was written mainly by Klein. Nicholls <strong>and</strong> Mimura contributed the UK <strong>and</strong><br />

Japanese case studies.<br />

This paper was written by Klein, based on the IPCC chapter by Klein et al. (2000). Klein<br />

was the co-ordinating lead author of the IPCC chapter, which benefited from text inputs<br />

from the other authors.<br />

Table 5 — Respective contributions by the PhD c<strong>and</strong>idate <strong>to</strong> the six published papers in this<br />

thesis.<br />

References<br />

Adger, W.N., 1997: Sustainability <strong>and</strong> Social Resilience in <strong>Coastal</strong> Resource Use. CSERGE Working Paper Series No.<br />

97-23, Centre for Social <strong>and</strong> Economic Research on the Global Environment, University of East Anglia, Norwich<br />

<strong>and</strong> University College London, UK, 39 pp.<br />

Adger, W.N., 2000: Social <strong>and</strong> ecological <strong>resilience</strong>: are they related? Progress in Human Geography, 24(3), 347–364.<br />

Adger, W.N. <strong>and</strong> P.M. Kelly, 1999: Social <strong>vulnerability</strong> <strong>to</strong> climate change <strong>and</strong> the architecture of entitlements. Mitigation<br />

<strong>and</strong> Adaptation Strategies for Global Change, 4(3–4), 253–266.<br />

Alverson, D.L., M.H. Freeborg, S.A. Murawski <strong>and</strong> J.A. Pope, 1994: A Global Assessment of Fisheries Bycatch <strong>and</strong> Discards.<br />

FAO Fisheries Technical Paper No. 339, Food <strong>and</strong> Agricultural Organisation, Rome, Italy, 233 pp.<br />

Anderson, D.M. <strong>and</strong> D.J. Garrison (eds.), 1997: The ecology <strong>and</strong> oceanography of harmful algal blooms. Limnology<br />

<strong>and</strong> Oceanography, 42(5,2), 1009–1305.<br />

Andreadakis, A.D., 1997: Wastewater treatment <strong>and</strong> disposal for the preservation of bathing <strong>and</strong> coastal water quality<br />

in <strong>to</strong>uristic areas. Marine Chemistry, 58(3–4), 389–395.<br />

Baarse, G., 1995: Development of an Operational Tool for Global Vulnerability Assessment (GVA)—Update of the<br />

Number of People at Risk Due <strong>to</strong> Sea Level Rise <strong>and</strong> Increased Flooding Probability. CZM Centre Publication No.<br />

3, Ministry of Transport, Public Works <strong>and</strong> Water Management, The Hague, The Netherl<strong>and</strong>s, 17 pp.<br />

Bijlsma, L., C.N. Ehler, R.J.T. Klein, S.M. Kulshrestha, R.F. McLean, N. Mimura, R.J. Nicholls, L.A. Nurse, H. Pérez<br />

Nie<strong>to</strong>, E.Z. Stakhiv, R.K. Turner <strong>and</strong> R.A. Warrick, 1996: <strong>Coastal</strong> zones <strong>and</strong> small isl<strong>and</strong>s. In: Climate Change<br />

1995—Impacts, Adaptations <strong>and</strong> Mitigation of Climate Change: Scientific-Technical Analyses, R.T. Watson, M.C.<br />

Zinyowera <strong>and</strong> R.H. Moss (eds.), Contribution of Working Group II <strong>to</strong> the Second Assessment Report of the Intergovernmental<br />

Panel on Climate Change, Cambridge University Press, Cambridge, UK, pp. 289–324.<br />

Briassoulis, H. <strong>and</strong> J. van der Straaten (eds.), 1992: Tourism <strong>and</strong> the Environment—Regional, Economic <strong>and</strong> Policy Issues.<br />

Environment & Assessment Series No. 2, Kluwer Academic Publishers, Dordrecht, The Netherl<strong>and</strong>s, 169<br />

pp.<br />

Brochier, F. <strong>and</strong> E. Ramieri, 2001: Climate Change Impacts on the Mediterranean <strong>Coastal</strong> Zones. Nota di Lavoro<br />

27.2001, Fondazione Eni Enrico Mattei, Milano, Italy, 82 pp.<br />

Bur<strong>to</strong>n, I., 1996: The growth of <strong>adaptation</strong> capacity: practice <strong>and</strong> policy. In: Adapting <strong>to</strong> Climate Change: An International<br />

Perspective, J.B. Smith, N. Bhatti, G.V. Menzhulin, R. Benioff, M. Campos, B. Jallow, F. Rijsberman,<br />

M.I. Budyko <strong>and</strong> R.K. Dixon (eds.), Springer-Verlag, New York, NY, USA, pp. 55–67.<br />

Bur<strong>to</strong>n, I., 1997: Vulnerability <strong>and</strong> adaptive response in the context of climate <strong>and</strong> climate change. Climatic Change,<br />

36(1–2), 185–196.<br />

Carpenter, S., B. Walker, J.M. Anderies <strong>and</strong> N. Abel, 2001: From metaphor <strong>to</strong> measurement: <strong>resilience</strong> of what <strong>to</strong><br />

what? Ecosystems, 4(8), 765–781.<br />

Carson, R.T., N.E. Flores, K.M. Martin <strong>and</strong> J.L. Wright, 1996: Contingent valuation <strong>and</strong> revealed preference methodologies:<br />

comparing the estimates for quasi-public goods. L<strong>and</strong> Economics, 72(1), 80–99.<br />

Carter, R.W.G., 1988: <strong>Coastal</strong> Environments—An Introduction <strong>to</strong> the Physical, Ecological <strong>and</strong> Cultural Systems of<br />

Coastlines. Academic Press, London, UK, xv+617 pp.<br />

Carter, T.R., M.L. Parry, S. Nishioka <strong>and</strong> H. Harasawa (eds.), 1994: Technical Guidelines for Assessing Climate<br />

Change Impacts <strong>and</strong> Adaptations. Report of Working Group II of the Intergovernmental Panel on Climate<br />

Change, University College London <strong>and</strong> Centre for Global Environmental Research, London, UK <strong>and</strong> Tsukuba,<br />

Japan, x+59 pp.<br />

Chapin, F.S., E.S. Zavaleta, V.Y. Eviner, R.L. Naylor, P.M. Vi<strong>to</strong>usek, H.L. Reynolds, D.U. Hooper, S. Lavorel, O.E.<br />

Sala, S.E. Hobbie, M.C. Mack <strong>and</strong> S. Diaz, 2000: Consequences of changing biodiversity. Nature, 405, 234–242.<br />

Cicin-Sain, B. (ed.), 1993: Integrated coastal management. Ocean & <strong>Coastal</strong> Management, 21(1–3), 1–352.<br />

26


Cicin-Sain, B. <strong>and</strong> R.W. Knecht, 1998: Integrated <strong>Coastal</strong> <strong>and</strong> Ocean Management: Concepts <strong>and</strong> Practices. Isl<strong>and</strong><br />

Press, Washing<strong>to</strong>n, DC, USA, 499 pp.<br />

Cicin-Sain, B., R.W. Knecht <strong>and</strong> G.W. Fisk, 1995: Growth in capacity for integrated coastal management since<br />

UNCED: an international perspective. Ocean & <strong>Coastal</strong> Management, 29(1–3), 93–123.<br />

Church, J.A., J.M. Gregory, P. Huybrechts, M. Kuhn, K. Lambeck, M.T. Nhuan, D. Qin <strong>and</strong> P.L. Woodworth, 2001:<br />

Changes in sea level. In: Climate Change 2001: The Scientific Basis, J.T. Hough<strong>to</strong>n, Y. Ding, D.J. Griggs, M. Noguer,<br />

P.J. van der Linden, X. Dai, K. Maskell <strong>and</strong> C.A. Johnson (eds.), Contribution of Working Group I <strong>to</strong> the<br />

Third Assessment Report of the Intergovernmental Panel on Climate Change, Cambridge University Press, Cambridge,<br />

UK, pp. 639–693.<br />

Cohen, J.E., C. Small, A. Mellinger, J. Gallup <strong>and</strong> J. Sachs, 1997: Estimates of coastal populations. Science, 278,<br />

1211–1212.<br />

Common, M.S., 1995: Sustainability <strong>and</strong> Policy: Limits <strong>to</strong> Economics. Cambridge University Press, Cambridge, UK,<br />

xii+348 pp.<br />

Costanza, R., R. d'Arge, R. de Groot, S. Farber, M. Grasso, B. Hannon, K. Limburg, S. Naeem, R. O'Neill, J. Paruelo,<br />

R. Raskin, P. Sut<strong>to</strong>n <strong>and</strong> M. van den Belt, 1997: The value of the world's ecosystem services <strong>and</strong> natural capital.<br />

Nature, 387, 253–260.<br />

De Groot, R.S., 1992a: Functions of Nature—Evaluation of Nature in Environmental Planning, Management <strong>and</strong> Decision<br />

Making. Wolters-Noordhoff, Groningen, The Netherl<strong>and</strong>s, xviii+315 pp.<br />

De Groot, R.S., 1992b: Functions <strong>and</strong> Economic Values of <strong>Coastal</strong> Protected Areas. MEDPAN Workshop on Economic<br />

Impact of the Mediterranean <strong>Coastal</strong> Protected Areas, Ajaccio, France, 26–28 September 1991, 17 pp.<br />

De la Vega-Leinert, A.C., R.J. Nicholls <strong>and</strong> R.S.J. Tol (eds.), 2000a: Proceedings of the SURVAS Expert Workshop on<br />

European Vulnerability <strong>and</strong> Adaptation <strong>to</strong> Accelerated Sea-Level Rise. Hamburg, Germany, 19–21 June 2000,<br />

Flood Hazard Research Centre, Middlesex University, Enfield, UK, viii+152 pp.<br />

De la Vega-Leinert, A.C., R.J. Nicholls, A. Nasser Hassan <strong>and</strong> M. El-Raey (eds.), 2000b: Proceedings of the SURVAS<br />

Expert Workshop on African Vulnerability <strong>and</strong> Adaptation <strong>to</strong> Accelerated Sea-Level Rise. Cairo, Egypt, 5–8 November<br />

2000, Flood Hazard Research Centre, Middlesex University, Enfield, UK, vi+104 pp.<br />

Demora, S.J. <strong>and</strong> E. Pelletier, 1997: Environmental tributyltin research—past, present, future. Environmental Technology,<br />

18(12), 1169–1177.<br />

Desenclos, J.C., 1996: Epidémiologie des risques <strong>to</strong>xiques et infectieux liés à la consommation de coquillages. Revue<br />

d’Epidémiologie et de Santé Publique, 44(5), 437–454.<br />

Dixon, J.A., L. Fallon Scura, R.A. Carpenter <strong>and</strong> P.B. Sherman, 1994: Economic Analysis of Environmental Impacts.<br />

Second edition. Earthscan Publications, London, UK, xii+210 pp.<br />

Dovers, S.R. <strong>and</strong> J.W. H<strong>and</strong>mer, 1992: Uncertainty, sustainability <strong>and</strong> change. Global Environmental Change, 2(4),<br />

262–276.<br />

Downing, T.E., M.J. Watts <strong>and</strong> H.G. Bohle, 1996: Climate change <strong>and</strong> food insecurity: <strong>to</strong>wards a sociology <strong>and</strong> geography<br />

of <strong>vulnerability</strong>. In: Climate Change <strong>and</strong> World Food Security, T.E. Downing (ed.), Springer Verlag, Berlin,<br />

Germany, pp. 183–206.<br />

Dulvy, N.K., D.P. Stanwell-Smith, W.R.T. Darwall <strong>and</strong> C.J. Horrill, 1995: Coral mining at Mafia Isl<strong>and</strong>, Tanzania: a<br />

management dilemma. Ambio, 24(6), 358–365.<br />

Eding<strong>to</strong>n, J.M. <strong>and</strong> M.A. Eding<strong>to</strong>n, 1986: Ecology, Recreation <strong>and</strong> Tourism. Cambridge University Press, Cambridge,<br />

UK, viii+200 pp.<br />

Ehler, C.N., 1993: Prepara<strong>to</strong>ry Workshop on Integrated <strong>Coastal</strong> Zone Management <strong>and</strong> Responses <strong>to</strong> Climate Change.<br />

Proceedings of the IPCC/WCC'93 Western Hemisphere Workshop, New Orleans, LA, USA, 13–16 July 1993, National<br />

Oceanic <strong>and</strong> Atmospheric Administration, Silver Spring, MD, USA, 611 pp.<br />

Ehler, C.N., B. Cicin-Sain, R. Knecht, R. South <strong>and</strong> R. Weiher, 1997: Guidelines <strong>to</strong> assist policy makers <strong>and</strong> managers<br />

of coastal areas in the integration of coastal management programs <strong>and</strong> national climate-change action plans.<br />

Ocean & <strong>Coastal</strong> Management, 37(1), 7–27.<br />

Emery, K.O. <strong>and</strong> D.G. Aubrey, 1991: Sea Levels, L<strong>and</strong> Levels, <strong>and</strong> Tide Gauges. Springer Verlag, New York, NY, USA,<br />

xiv+237 pp.<br />

Fankhauser, S., 1995: Protection vs. retreat: estimating the costs of sea level rise. Environment <strong>and</strong> Planning A,<br />

27(2), 299–319.<br />

Feenstra, J.F., I. Bur<strong>to</strong>n, J.B. Smith <strong>and</strong> R.S.J. Tol (eds.), 1998: H<strong>and</strong>book on Methods for Climate Change Impact<br />

Assessment <strong>and</strong> Adaptation Strategies. Version 2.0, United Nations Environment Programme <strong>and</strong> Institute for<br />

Environmental Studies, Vrije Universiteit, Nairobi, Kenya <strong>and</strong> Amsterdam, The Netherl<strong>and</strong>s, xxvi+422 pp.<br />

Folke, C., S. Carpenter, T. Elmqvist, L. Gunderson, C.S. Holling, B. Walker, J. Bengtsson, F. Berkes, J. Colding, K.<br />

Danell, M. Falkenmark, L. Gordon, R. Kasperson, N. Kautsky, A. Kinzig, S. Levin, K.-G. Mäler, F. Moberg, L.<br />

Ohlsson, P. Olsson, E. Ostrom, W. Reid, J. Rockstroem, H. Savenije <strong>and</strong> U. Svedin, 2002: Resilience <strong>and</strong> Sustainable<br />

Development: Building Adaptive Capacity in a World of Transformations. Environmental Advisory<br />

Council, S<strong>to</strong>ckholm, Sweden, 73 pp.<br />

Gilbert, A.J. <strong>and</strong> R. Janssen, 1998: Use of environmental functions <strong>to</strong> communicate the values of a mangrove ecosystem<br />

under different management regimes. Ecological Economics, 25(3), 323–346.<br />

Goldberg, E.D., 1994: <strong>Coastal</strong> Zone Space—Prelude <strong>to</strong> Conflict? IOC Ocean Forum I, UNESCO Publishing, Paris,<br />

France, 138 pp.<br />

Gommes, R., J. du Guerny, F. Nachtergaele <strong>and</strong> R. Brinkman, 1998: Potential Impacts of Sea-Level Rise on Populations<br />

<strong>and</strong> Agriculture. FAO SD-Dimensions Special, Food <strong>and</strong> Agricultural Organisation, Rome, Italy,<br />

http://www.fao.org/sd/eidirect/EIre0045.htm.<br />

Gordon, J.E., 1978: Structures—Or Why Things Don’t Fall Down. Penguin Books, Harmondsworth, UK, 395 pp.<br />

Gornitz, V., 2000: Impoundment, groundwater mining, <strong>and</strong> other hydrologic transformations: impacts on global sea<br />

level rise. In: Sea Level Rise: His<strong>to</strong>ry <strong>and</strong> Consequences, B.C. Douglas, M.S. Kearney <strong>and</strong> S.P. Leatherman<br />

(eds.), Academic Press, San Diego, CA, USA, pp. 97–119.<br />

Gornitz, V.M., R.C. Daniels, T.W. White <strong>and</strong> K.R. Birdwell, 1994: The development of a coastal risk assessment database:<br />

<strong>vulnerability</strong> <strong>to</strong> sea-level rise in the U.S. Southeast. Journal of <strong>Coastal</strong> Research, special issue 12, 327–<br />

338.<br />

27


Gornitz, V., C. Rosenzweig <strong>and</strong> D. Hillel, 1997: Effects of anthropogenic intervention in the l<strong>and</strong> hydrological cycle<br />

on global sea level rise. Global <strong>and</strong> Planetary Change, 14(3–4), 147–161.<br />

Grainger, R.J.R., 1997: Recent Trends in Global Fishery Production. FAO Fisheries Department, Rome, Italy,<br />

http://www.fao.org/fi/trends/catch/catch.asp.<br />

Grainger, R.J.R. <strong>and</strong> S.M. Garcia, 1996: Chronicles of Marine Fishery L<strong>and</strong>ings (1950–1994): Trend Analysis <strong>and</strong> Fisheries<br />

Potential. FAO Fisheries Technical Paper No. 359, Food <strong>and</strong> Agricultural Organisation, Rome, Italy, 51 pp.<br />

Gren, I.-M., C. Folke, R.K. Turner <strong>and</strong> I.J. Bateman, 1994: Primary <strong>and</strong> secondary values of wetl<strong>and</strong> ecosystems. Environmental<br />

<strong>and</strong> Resource Economics, 4(1), 55–74.<br />

Gröger, M. <strong>and</strong> H.-P. Plag, 1993: Estimations of a global sea level trend: limitations from the structure of the PSMSL<br />

global sea level data set. Global <strong>and</strong> Planetary Change, 8(3), 161–179.<br />

Gull<strong>and</strong>, J.A. (ed.), 1971: The Fish Resources of the Ocean. Fishing News, West Byfleet, UK, xxvi+255 pp.<br />

Hanley, N. <strong>and</strong> C.L. Spash, 1993: Cost-Benefit Analysis <strong>and</strong> the Environment. Edward Elgar, Aldershot, UK, x+278 pp.<br />

H<strong>and</strong>mer, J.W. <strong>and</strong> S.R. Dovers, 1996: A typology of <strong>resilience</strong>: rethinking institutions for sustainable development.<br />

Industrial <strong>and</strong> Environmental Crisis Quarterly, 9(4), 482–511.<br />

Harvey, N., B. Clous<strong>to</strong>n <strong>and</strong> P. Carvalho, 1999: Improving coastal <strong>vulnerability</strong> assessment methodologies for integrated<br />

coastal zone management: an approach from South Australia. Australian Geographical Studies, 37(1),<br />

50–69.<br />

Hawkins, J.P. <strong>and</strong> C.M. Roberts, 1994: The growth of coastal <strong>to</strong>urism in the Red Sea: present <strong>and</strong> future effects on<br />

coral reefs. Ambio, 23(8), 503–508.<br />

Holling, C.S., 1973: Resilience <strong>and</strong> stability of ecological systems. Annual Review of Ecology <strong>and</strong> Systematics, 4, 1–<br />

23.<br />

Hoozemans, F.M.J., M. March<strong>and</strong> <strong>and</strong> H.A. Pennekamp, 1993: Sea Level Rise: A Global Vulnerability Assessment—<br />

Vulnerability Assessments for Population, <strong>Coastal</strong> Wetl<strong>and</strong>s <strong>and</strong> Rice Production on a Global Scale. 2nd revised<br />

edition, Delft Hydraulics <strong>and</strong> Rijkswaterstaat, Delft <strong>and</strong> The Hague, The Netherl<strong>and</strong>s, xxxii+184 pp.<br />

Hough<strong>to</strong>n, T., Y. Ding, D.J. Griggs, M. Noguer, P.J. van der Linden, X. Dai, K. Maskell <strong>and</strong> C.A. Johnson (eds.), 2001:<br />

Climate Change 2001: The Scientific Basis. Contribution of Working Group I <strong>to</strong> the Third Assessment Report of<br />

the Intergovernmental Panel on Climate Change, Cambridge University Press, Cambridge, UK, x+881 pp.<br />

IPCC CZMS, 1990: Strategies for Adaptation <strong>to</strong> Sea Level Rise. Report of the <strong>Coastal</strong> Zone Management Subgroup, Response<br />

Strategies Working Group of the Intergovernmental Panel on Climate Change, Ministry of Transport,<br />

Public Works <strong>and</strong> Water Management, The Hague, The Netherl<strong>and</strong>s, x+122 pp.<br />

IPCC CZMS, 1992: A common methodology for assessing <strong>vulnerability</strong> <strong>to</strong> sea-level rise—second revision. In: Global<br />

Climate Change <strong>and</strong> the Rising Challenge of the Sea, Report of the <strong>Coastal</strong> Zone Management Subgroup, Response<br />

Strategies Working Group of the Intergovernmental Panel on Climate Change, Ministry of Transport,<br />

Public Works <strong>and</strong> Water Management, The Hague, The Netherl<strong>and</strong>s, Appendix C, 27 pp.<br />

ISDR, 2002: Living with Risk: A Global Review of Disaster Reduction Initiatives. Secretariat of the International<br />

Strategy for Disaster Reduction, Geneva, Switzerl<strong>and</strong>, iii+382 pp.<br />

Jeftić, L., J. Milliman <strong>and</strong> G. Sestini (eds.), 1992: Climate Change <strong>and</strong> the Mediterranean—Environmental <strong>and</strong> Societal<br />

Impacts of Climatic Change <strong>and</strong> Sea-Level Rise in the Mediterranean Region. Edward Arnold, London, UK,<br />

xii+673 pp.<br />

Jeftić, L., S. Kečkeš <strong>and</strong> J.C. Pernetta (eds.), 1996: Climate Change <strong>and</strong> the Mediterranean—Environmental <strong>and</strong> Societal<br />

Impacts of Climate Change <strong>and</strong> Sea Level Rise in the Mediterranean Region. Volume 2, Arnold, London,<br />

UK, xi+564 pp.<br />

Kates, R.W, 1997: Climate change 1995—impacts, <strong>adaptation</strong>s, <strong>and</strong> mitigation. Environment, 39(9), 29–33.<br />

Kay, R.C. <strong>and</strong> J. Alder, 1999: <strong>Coastal</strong> Planning <strong>and</strong> Management. E & FN Spon, London, UK, xxi+375 pp.<br />

Kay, R.C., I. Eliot, B. Ca<strong>to</strong>n, G. Morvell <strong>and</strong> P. Waterman, 1996: A review of the Intergovernmental Panel on Climate<br />

Change’s Common Methodology for Assessing the Vulnerability of <strong>Coastal</strong> Areas <strong>to</strong> Sea-Level Rise. <strong>Coastal</strong> Management,<br />

24, 165–188.<br />

Kay, R.C. <strong>and</strong> J.E. Hay, 1993: A decision support approach <strong>to</strong> coastal <strong>vulnerability</strong> <strong>and</strong> <strong>resilience</strong> assessment: a <strong>to</strong>ol<br />

for integrated coastal zone management. In: Vulnerability Assessment <strong>to</strong> Sea-Level Rise <strong>and</strong> <strong>Coastal</strong> Zone Management,<br />

R.F. McLean <strong>and</strong> N. Mimura (eds.), Proceedings of the IPCC/WCC’93 Eastern Hemisphere Workshop,<br />

Tsukuba, Japan, 3–6 August 1993, Department of Environment, Sport <strong>and</strong> Terri<strong>to</strong>ries, Canberra, Australia, pp.<br />

213–225.<br />

Klein, R.J.T., 1998: Towards better underst<strong>and</strong>ing, assessment <strong>and</strong> funding of climate <strong>adaptation</strong>. Change, 44, 15–<br />

19.<br />

Klein, R.J.T., 2001: Adaptation <strong>to</strong> Climate Change in German Official Development Assistance—An Inven<strong>to</strong>ry of Activities<br />

<strong>and</strong> Opportunities, with a Special Focus on Africa. Deutsche Gesellschaft für Technische Zusammenarbeit,<br />

Eschborn, Germany, 42 pp.<br />

Klein, R.J.T. <strong>and</strong> I.J. Bateman, 1998: The recreational value of Cley Marshes Nature Reserve: an argument against<br />

managed retreat? Water <strong>and</strong> Environmental Management, 12(4), 280–285.<br />

Klein, R.J.T., E.N. Buckley, J. As<strong>to</strong>n, R.J. Nicholls, S. Ragoonaden, M. Capobianco, N. Mizutani <strong>and</strong> P.D. Nunn, 2000:<br />

<strong>Coastal</strong> <strong>adaptation</strong>. In: Methodological <strong>and</strong> Technological Issues in Technology Transfer, B. Metz, O.R. Davidson,<br />

J.W. Martens, S.N.M. van Rooijen <strong>and</strong> L.L. Van Wie-McGrory (eds.), Special Report of Working Group III of<br />

the Intergovernmental Panel on Climate Change, Cambridge University Press, Cambridge, UK, pp. 349–372.<br />

Klein, R.J.T. <strong>and</strong> D.C. MacIver, 1999: Adaptation <strong>to</strong> climate variability <strong>and</strong> change: methodological issues. Mitigation<br />

<strong>and</strong> Adaptation Strategies for Global Change, 4(3–4), 189–198<br />

Klein, R.J.T. <strong>and</strong> R.J. Nicholls, 1998: <strong>Coastal</strong> zones. In: H<strong>and</strong>book on Methods for Climate Change Impact Assessment<br />

<strong>and</strong> Adaptation Strategies, J.F. Feenstra, I. Bur<strong>to</strong>n, J.B. Smith <strong>and</strong> R.S.J. Tol (eds.), Version 2.0, United Nations<br />

Environment Programme <strong>and</strong> Institute for Environmental Studies, Vrije Universiteit, Nairobi, Kenya <strong>and</strong><br />

Amsterdam, The Netherl<strong>and</strong>s, pp. 7.1–7.35.<br />

Klein, R.J.T. <strong>and</strong> R.J. Nicholls, 1999: Assessment of coastal <strong>vulnerability</strong> <strong>to</strong> climate change. Ambio, 28(2), 182–187.<br />

Klein, R.J.T., R.J. Nicholls <strong>and</strong> N. Mimura, 1999: <strong>Coastal</strong> <strong>adaptation</strong> <strong>to</strong> climate change: can the IPCC Technical<br />

Guidelines be applied? Mitigation <strong>and</strong> Adaptation Strategies for Global Change, 4(3–4), 51–64.<br />

28


Klein, R.J.T., R.J. Nicholls, S. Ragoonaden, M. Capobianco, J. As<strong>to</strong>n <strong>and</strong> E.N. Buckley, 2001: Technological options<br />

for <strong>adaptation</strong> <strong>to</strong> climate change in coastal zones. Journal for <strong>Coastal</strong> Research, 17(3), 531–543.<br />

Klein, R.J.T., R.J. Nicholls <strong>and</strong> F. Thomalla, 2002: The Resilience of <strong>Coastal</strong> Megacities <strong>to</strong> Weather-Related Hazards.<br />

Workshop "The Future of Disaster Risk: Building Safer Cities", Disaster Management Facility, World Bank, Washing<strong>to</strong>n,<br />

DC, USA, 4–6 December 2002, 21 pp.<br />

Klein, R.J.T., M.J. Smit, H. Goosen <strong>and</strong> C.H. Hulsbergen, 1998: Resilience <strong>and</strong> <strong>vulnerability</strong>: coastal dynamics or<br />

Dutch dikes? The Geographical Journal, 164(3), 259–268.<br />

Klein, R.J.T. <strong>and</strong> R.S.J. Tol, 1997: Adaptation <strong>to</strong> Climate Change: Options <strong>and</strong> Technologies—An Overview Paper.<br />

Technical Paper FCCC/TP/1997/3, United Nations Framework Convention on Climate Change Secretariat, Bonn,<br />

Germany, 33 pp.<br />

Knight, D., B. Mitchell <strong>and</strong> G. Wall, 1997: Bali: sustainable development, <strong>to</strong>urism <strong>and</strong> coastal management. Ambio,<br />

26(2), 90–96.<br />

Leary, N.A., 1999: A framework for benefit-cost analysis of <strong>adaptation</strong> <strong>to</strong> climate change <strong>and</strong> climate variability,<br />

Mitigation <strong>and</strong> Adaptation Strategies for Global Change, 4(3–4), 307–318.<br />

Leatherman, S.P. (ed.), 1997: Isl<strong>and</strong> states at risk: global climate change, development <strong>and</strong> population. Journal of<br />

<strong>Coastal</strong> Research, special issue 24, 1–239.<br />

Lenhart, S., S. Huq, L. José Mata, I. Nemešová, S. Toure <strong>and</strong> J.B. Smith (eds.), 1996: Vulnerability <strong>and</strong> Adaptation<br />

<strong>to</strong> Climate Change—A Synthesis of Results from the U.S. Country Studies Program. Interim Report, U.S. Country<br />

Studies Program, Washing<strong>to</strong>n, DC, USA, xxxv+366 pp.<br />

Maul, G.A. (ed.), 1993: Climatic Change in the Intra-Americas Sea—Implications of Future Climate on the Ecosystems<br />

<strong>and</strong> Socio-Economic Structure in the Marine <strong>and</strong> <strong>Coastal</strong> Regions of the Caribbean Sea, Gulf of Mexico, Bahamas,<br />

<strong>and</strong> the Northeast Coast of South America. Edward Arnold, London, UK, xii+389 pp.<br />

McLean, R.F. <strong>and</strong> N. Mimura (eds.), 1993: Vulnerability Assessment <strong>to</strong> Sea Level Rise <strong>and</strong> <strong>Coastal</strong> Zone Management.<br />

Proceedings of the IPCC/WCC'93 Eastern Hemisphere Workshop, Tsukuba, Japan, 3–6 August 1993, Department<br />

of Environment, Sport <strong>and</strong> Terri<strong>to</strong>ries, Canberra, Australia, viii+429 pp.+apps.<br />

Miller, M.L. <strong>and</strong> J. Auyong, 1991: <strong>Coastal</strong> zone <strong>to</strong>urism—a potent force affecting environment <strong>and</strong> society. Marine<br />

Policy, March, 75–99.<br />

Mimura, N., M. Isobe <strong>and</strong> Y. Hosakawa, 1992: Impacts of sea level rise on Japanese coastal zones <strong>and</strong> response<br />

strategies. In: Global Climate Change <strong>and</strong> the Rising Challenge of the Sea, J. O'Callahan, (ed.). Proceedings of<br />

the Third IPCC CZMS Workshop, Margarita Isl<strong>and</strong>, Venezuela, 9–13 March 1992, National Oceanic <strong>and</strong> Atmospheric<br />

Administration, Silver Spring, MD, USA, pp. 329–349.<br />

Mimura, N. <strong>and</strong> H. Yokoki (eds.), 2001: Global Change <strong>and</strong> Asia Pacific Coasts. Proceedings of the APN-SURVAS-LOICZ<br />

Joint Conference on <strong>Coastal</strong> Impacts of Climate Change <strong>and</strong> Adaptation in the Asia-Pacific Region, Kobe, Japan,<br />

14–16 November 2000, Asia Pacific Network for Global Change Research <strong>and</strong> Ibaraki University, Ibaraki, Japan,<br />

x+285 pp.<br />

Nakićenović, N. <strong>and</strong> R. Swart (eds.), 2000: Emission Scenarios. Special Report of Working Group III of the Intergovernmental<br />

Panel on Climate Change, Cambridge University Press, Cambridge, UK, 599 pp.<br />

Nelson, J.G., R. Butler <strong>and</strong> G. Wall (eds.), 1993: Tourism <strong>and</strong> Sustainable Development: Moni<strong>to</strong>ring, Planning, Managing.<br />

Department of Geography Publication Series No. 37, Heritage Resources Centre Joint Publication No. 1,<br />

University of Waterloo, Waterloo, Canada, xxii+283 pp.<br />

Nicholls, R.J., 1995a: <strong>Coastal</strong> megacities <strong>and</strong> climate change. GeoJournal, 37(3), 369–379.<br />

Nicholls, R.J., 1995b: Synthesis of <strong>vulnerability</strong> analysis studies. In: Proceedings of the World Coast Conference 1993,<br />

P. Beukenkamp, P. Günther, R.J.T. Klein, R. Misdorp, D. Sadacharan <strong>and</strong> L.P.M. de Vrees (eds.), Noordwijk,<br />

The Netherl<strong>and</strong>s, 1–5 November 1993, <strong>Coastal</strong> Zone Management Centre Publication No. 4, National Institute<br />

for <strong>Coastal</strong> <strong>and</strong> Marine Management, The Hague, The Netherl<strong>and</strong>s, pp. 181–216.<br />

Nicholls, R.J., 2002: Analysis of global impacts of sea-level rise: a case study of flooding. Physics <strong>and</strong> Chemistry of<br />

the Earth, Parts A/B/C, 27(32–34), 1455–1466.<br />

Nicholls, R.J. <strong>and</strong> J. Branson (eds.), 1998: Enhancing coastal <strong>resilience</strong>—planning for an uncertain future. The Geographical<br />

Journal, 164(3), 255–343.<br />

Nicholls, R.J., F.M.J. Hoozemans <strong>and</strong> M. March<strong>and</strong>, 1999: Increasing flood risk <strong>and</strong> wetl<strong>and</strong> losses due <strong>to</strong> global sealevel<br />

rise: regional <strong>and</strong> global analyses—the science of climate change. Global Environmental Change, 9, S69–<br />

S87.<br />

Nicholls, R.J. <strong>and</strong> S.P. Leatherman (eds.), 1995: The potential impact of accelerated sea-level rise on developing<br />

countries. Journal of <strong>Coastal</strong> Research, special issue 14, 1–324.<br />

Nicholls, R.J. <strong>and</strong> N. Mimura, 1998: Regional issues raised by sea-level rise <strong>and</strong> their policy implications. Climate Research,<br />

11(1), 5–18.<br />

Nicholls, R.J. <strong>and</strong> C. Small, 2002: Improved estimates of coastal population <strong>and</strong> exposure <strong>to</strong> hazards released. EOS<br />

Transactions, 83(2), 301, 305.<br />

O’Callahan, J. (ed.), 1994: Global Climate Change <strong>and</strong> the Rising Challenge of the Sea. Proceedings of the third IPCC<br />

CZMS Workshop, Isla de Margarita, Venezuela, 9–13 March 1992, National Oceanic <strong>and</strong> Atmospheric Administration,<br />

Silver Spring, MD, USA, v+691 pp.<br />

OECD, 1997: Tourism Policy <strong>and</strong> International Tourism in OECD Countries. Publication No. OCDE/GD(97)173, Organisation<br />

for Economic Co-operation <strong>and</strong> Development, Paris, France, 184 pp.<br />

Parry, M., N. Arnell, M. Hulme, R. Nicholls <strong>and</strong> M. Livermore, 1998: Adapting <strong>to</strong> the inevitable. Nature, 395, 741.<br />

Pearce, D.W. <strong>and</strong> R.K. Turner, 1990: Economics of Natural Resources <strong>and</strong> the Environment. Harvester Wheatsheaf,<br />

Hemel Hempstead, UK, xiii+378 pp.<br />

Peterson, G.D., C.R. Allen <strong>and</strong> C.S. Holling, 1998: Ecological <strong>resilience</strong>, biodiversity <strong>and</strong> scale. Ecosystems, 1(1), 6–<br />

18.<br />

Pielke, R.A., Jr., 1998: Rethinking the role of <strong>adaptation</strong> in climate policy. Global Environmental Change, 8(2), 159–<br />

170.<br />

Pimm, S.L., 1984: The complexity <strong>and</strong> stability of ecosystems. Nature, 307, 321–326.<br />

Post, J.C. <strong>and</strong> C.G. Lundin, 1996: Guidelines for Integrated <strong>Coastal</strong> Zone Management. Environmentally Sustainable<br />

Development Studies <strong>and</strong> Monographs Series No. 9, The World Bank, Washing<strong>to</strong>n, DC, USA, 16 pp.<br />

29


Pruss, A., 1998: Review of epidemiological studies on health effects from exposure <strong>to</strong> recreational water. International<br />

Journal of Epidemiology, 27(1), 1–9.<br />

Sahagian, D., 2000: Global physical effects of anthropogenic hydrological alterations: sea level <strong>and</strong> water redistribution.<br />

Global <strong>and</strong> Planetary Change, 25(1–2), 39–48.<br />

Sahagian, D.L., F.W. Schwartz <strong>and</strong> D.K. Jacobs, 1994: Direct anthropogenic contributions <strong>to</strong> sea level rise in the<br />

twentieth century. Nature, 367, 54–57.<br />

Schulze, E.D. <strong>and</strong> H.A. Mooney, 1993: Biodiversity <strong>and</strong> Ecosystem Function. Springer-Verlag, Berlin, Germany,<br />

xxvii+525 pp.<br />

Small, C. <strong>and</strong> J.E. Cohen, 1999: Continental physiography, climate <strong>and</strong> the global distribution of human population.<br />

In: Toward Digital Earth, G. Xu <strong>and</strong> Y. Chen (eds.), Proceedings of the International Symposium on Digital<br />

Earth, Beijing, China, 29 November – 2 December 1999, Science Press, Beijing, China, pp. 965–971.<br />

Smit, B., I. Bur<strong>to</strong>n, R.J.T. Klein <strong>and</strong> J. W<strong>and</strong>el, 2000: An ana<strong>to</strong>my of <strong>adaptation</strong> <strong>to</strong> climate change <strong>and</strong> variability.<br />

Climatic Change, 45(1), 223–251.<br />

Smith, K., 1992: Environmental Hazards: Assessing Risk <strong>and</strong> Reducing Disaster. Routledge Physical Environment Series,<br />

Routledge, London, UK, xx+324 pp.<br />

Ten Hallers-Tjabbes, C.C., 1997: Tributyltin <strong>and</strong> policies for antifouling. Environmental Technology, 18(12), 1265–<br />

1268.<br />

Timmerman, P., 1981: Vulnerability, Resilience <strong>and</strong> the Collapse of Society: A Review of Models <strong>and</strong> Possible Climatic<br />

Applications. Environmental Monograph No. 1, Institute of Environmental Studies, University of Toron<strong>to</strong>,<br />

Toron<strong>to</strong>, Canada, ii+42 pp.<br />

Tol, R.S.J., R.J.T. Klein, H.M.A. Jansen <strong>and</strong> H. Verbruggen, 1996: Some economic considerations on the importance<br />

of proactive integrated coastal zone management. Ocean & <strong>Coastal</strong> Management, 32(1), 39–55.<br />

Tooley, M. <strong>and</strong> S. Jelgersma (eds.), 1992: Impacts of Sea Level Rise on European <strong>Coastal</strong> Lowl<strong>and</strong>s. Blackwell, Oxford,<br />

UK, xv+267 pp.<br />

Turner, R.K., 1988: Wetl<strong>and</strong> conservation: economics <strong>and</strong> ethics. In: Economics, Growth <strong>and</strong> Sustainable Environments—Essays<br />

in Memory of Richard Lecomber, D. Collard, D.W. Pearce <strong>and</strong> D. Ulph (eds.), Macmillan,<br />

Houndsmills, UK, pp. 121–159.<br />

Turner, R.K. <strong>and</strong> W.N. Adger, 1996: <strong>Coastal</strong> Zone Resources Assessment Guidelines. LOICZ Reports & Studies No. 4,<br />

L<strong>and</strong>-Ocean Interactions in the <strong>Coastal</strong> Zone, Texel, The Netherl<strong>and</strong>s, iv+101 pp.<br />

Turner, R.K., S. Subak <strong>and</strong> W.N. Adger, 1996: Pressures, trends, <strong>and</strong> impacts in coastal zones: interactions between<br />

socioeconomic <strong>and</strong> natural systems. Environmental Management, 20(2), 159–173.<br />

UNPD, 1999: The World at Six Billion. Department of Economic <strong>and</strong> Social Affairs, United Nations Secretariat, New<br />

York, NY, USA, vii+63 pp.<br />

UNPD, 2001: World Population Prospects—The 2000 Revision. Department of Economic <strong>and</strong> Social Affairs, United Nations<br />

Secretariat, New York, NY, USA, xii+64 pp.<br />

UNEP, 2002: Global Environment Outlook 3—Past, Present <strong>and</strong> Future Perspectives. Earthscan, London, UK,<br />

xxxiii+426 pp.<br />

Vallega, A., 1996: The coastal use structure within the coastal system—a sustainable development-consistent approach.<br />

Journal of Marine Systems, 7(1), 95–115.<br />

Vellinga, P., R.S. de Groot <strong>and</strong> R.J.T. Klein, 1994: An ecologically sustainable biosphere. In: The Environment: Towards<br />

a Sustainable Future, Dutch Committee for Long-Term Environmental Policy (ed.), Kluwer Academic<br />

Publishers, Dordrecht, The Netherl<strong>and</strong>s, pp. 317–346.<br />

Vellinga, P. <strong>and</strong> S.P. Leatherman, 1989: Sea level rise, consequences <strong>and</strong> policies. Climatic Change, 15(1–2), 175–<br />

189.<br />

Vörösmarty, C.J. <strong>and</strong> D. Sahagian, 2000: Anthropogenic disturbance of the terrestrial water cycle. Bioscience, 50(9),<br />

753–765.<br />

Watson, R.T., M.C. Zinyowera, R.H. Moss <strong>and</strong> D.J. Dokken (eds.), 1998: The Regional Impacts of Climate Change—An<br />

Assessment of Vulnerability. Special Report of the Intergovernmental Panel on Climate Change, Cambridge<br />

University Press, Cambridge, UK, x+517 pp.<br />

WCC’93, 1994: Preparing <strong>to</strong> Meet the <strong>Coastal</strong> Challenges of the 21st Century. Report of the World Coast Conference,<br />

Noordwijk, The Netherl<strong>and</strong>s, 1–5 November 1993. Ministry of Transport, Public Works <strong>and</strong> Water Management,<br />

The Hague, The Netherl<strong>and</strong>s, 49 pp.+apps.<br />

West, J.J. <strong>and</strong> H. Dowlatabadi, 1999: On assessing the economic impacts of sea-level rise on developed coasts. In:<br />

Climate, Change <strong>and</strong> Risk, T.E. Downing, A.A. Olsthoorn <strong>and</strong> R.S.J. Tol (eds.), Routledge, London, UK, pp. 205–<br />

220.<br />

West, J.J., M.J. Small <strong>and</strong> H. Dowlatabadi, 2001: S<strong>to</strong>rms, inves<strong>to</strong>r decisions <strong>and</strong> the economic impacts of sea level<br />

rise. Climatic Change, 48(2–3), 317–342.<br />

White, A.T., V. Barker <strong>and</strong> G. Tantrigama, 1997: Using integrated coastal management <strong>and</strong> economics <strong>to</strong> conserve<br />

coastal <strong>to</strong>urism resources in Sri Lanka. Ambio, 26(6), 335–344.<br />

WRI, 1996: World Resources 1996–97: A Guide <strong>to</strong> the Global Environment—The Urban Environment. World Resources<br />

Institute, United Nations Environment Programme, United Nations Development Programme <strong>and</strong> The World<br />

Bank, Oxford University Press, Oxford, UK, xvi+368 pp.<br />

Yohe, G.W. <strong>and</strong> J. Neumann, 1997: Planning for sea level rise <strong>and</strong> shore protection under climate uncertainty. Climatic<br />

Change, 37(1), 243–270.<br />

Yohe, G.W., J. Neumann, P. Marshall <strong>and</strong> H. Ameden, 1996: The economic cost of greenhouse-induced sea-level rise<br />

for developed property in the United States. Climatic Change, 32(3), 387–410.<br />

30

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!