30.01.2013 Views

River Restoration Managing the Uncertainty in Restoring ... - Inecol

River Restoration Managing the Uncertainty in Restoring ... - Inecol

River Restoration Managing the Uncertainty in Restoring ... - Inecol

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>River</strong> <strong>Restoration</strong><br />

<strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong><br />

Restor<strong>in</strong>g Physical Habitat<br />

Editors<br />

Stephen Darby<br />

School of Geography, University of Southampton, UK<br />

and<br />

David Sear<br />

School of Geography, University of Southampton, UK


<strong>River</strong> <strong>Restoration</strong>


<strong>River</strong> <strong>Restoration</strong><br />

<strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong><br />

Restor<strong>in</strong>g Physical Habitat<br />

Editors<br />

Stephen Darby<br />

School of Geography, University of Southampton, UK<br />

and<br />

David Sear<br />

School of Geography, University of Southampton, UK


Copyright © 2008 John Wiley & Sons Ltd, The Atrium, Sou<strong>the</strong>rn Gate, Chichester,<br />

West Sussex PO19 8SQ, England<br />

Telephone (+44) 1243 779777<br />

Email (for orders and customer service enquiries): cs-books@wiley.co.uk<br />

Visit our Home Page on www.wileyeurope.com or www.wiley.com<br />

All Rights Reserved. No part of this publication may be reproduced, stored <strong>in</strong> a retrieval system or transmitted <strong>in</strong> any form or by any<br />

means, electronic, mechanical, photocopy<strong>in</strong>g, record<strong>in</strong>g, scann<strong>in</strong>g or o<strong>the</strong>rwise, except under <strong>the</strong> terms of <strong>the</strong> Copyright, Designs and<br />

Patents Act 1988 or under <strong>the</strong> terms of a licence issued by <strong>the</strong> Copyright Licens<strong>in</strong>g Agency Ltd, 90 Tottenham Court Road, London W1T<br />

4LP, UK, without <strong>the</strong> permission <strong>in</strong> writ<strong>in</strong>g of <strong>the</strong> Publisher. Requests to <strong>the</strong> Publisher should be addressed to <strong>the</strong> Permissions Department,<br />

John Wiley & Sons Ltd, The Atrium, Sou<strong>the</strong>rn Gate, Chichester, West Sussex PO19 8SQ, England, or emailed to permreq@wiley.co.uk, or<br />

faxed to (+44) 1243 770620.<br />

Designations used by companies to dist<strong>in</strong>guish <strong>the</strong>ir products are often claimed as trademarks. All brand names and product names used <strong>in</strong><br />

this book are trade names, service marks, trademarks or registered trademarks of <strong>the</strong>ir respective owners. The Publisher is not associated<br />

with any product or vendor mentioned <strong>in</strong> this book.<br />

This publication is designed to provide accurate and authoritative <strong>in</strong>formation <strong>in</strong> regard to <strong>the</strong> subject matter covered. It is sold on <strong>the</strong><br />

understand<strong>in</strong>g that <strong>the</strong> Publisher is not engaged <strong>in</strong> render<strong>in</strong>g professional services. If professional advice or o<strong>the</strong>r expert assistance is<br />

required, <strong>the</strong> services of a competent professional should be sought.<br />

The Publisher and <strong>the</strong> Author make no representations or warranties with respect to <strong>the</strong> accuracy or completeness of <strong>the</strong> contents of this<br />

work and specifi cally disclaim all warranties, <strong>in</strong>clud<strong>in</strong>g without limitation any implied warranties of fi tness for a particular purpose. The<br />

advice and strategies conta<strong>in</strong>ed here<strong>in</strong> may not be suitable for every situation. In view of ongo<strong>in</strong>g research, equipment modifi cations,<br />

changes <strong>in</strong> governmental regulations, and <strong>the</strong> constant fl ow of <strong>in</strong>formation relat<strong>in</strong>g to <strong>the</strong> use of experimental reagents, equipment, and<br />

devices, <strong>the</strong> reader is urged to review and evaluate <strong>the</strong> <strong>in</strong>formation provided <strong>in</strong> <strong>the</strong> package <strong>in</strong>sert or <strong>in</strong>structions for each chemical, piece<br />

of equipment, reagent, or device for, among o<strong>the</strong>r th<strong>in</strong>gs, any changes <strong>in</strong> <strong>the</strong> <strong>in</strong>structions or <strong>in</strong>dication of usage and for added warn<strong>in</strong>gs and<br />

precautions. The fact that an organization or Website is referred to <strong>in</strong> this work as a citation and/or a potential source of fur<strong>the</strong>r <strong>in</strong>formation<br />

does not mean that <strong>the</strong> author or <strong>the</strong> publisher endorses <strong>the</strong> <strong>in</strong>formation <strong>the</strong> organization or Website may provide or recommendations it<br />

may make. Fur<strong>the</strong>r, readers should be aware that Internet Websites listed <strong>in</strong> this work may have changed or disappeared between when this<br />

work was written and when it is read. No warranty may be created or extended by any promotional statements for this work. Nei<strong>the</strong>r <strong>the</strong><br />

Publisher nor <strong>the</strong> Author shall be liable for any damages aris<strong>in</strong>g herefrom.<br />

O<strong>the</strong>r Wiley Editorial Offi ces<br />

John Wiley & Sons Inc., 111 <strong>River</strong> Street, Hoboken, NJ 07030, USA<br />

Jossey-Bass, 989 Market Street, San Francisco, CA 94103-1741, USA<br />

Wiley-VCH Verlag GmbH, Boschstr. 12, D-69469 We<strong>in</strong>heim, Germany<br />

John Wiley & Sons Australia Ltd, 42 McDougall Street, Milton, Queensland 4064, Australia<br />

John Wiley & Sons (Asia) Pte Ltd, 2 Clementi Loop #02-01, J<strong>in</strong> X<strong>in</strong>g Distripark, S<strong>in</strong>gapore 129809<br />

John Wiley & Sons Ltd, 6045 Freemont Blvd, Mississauga, Ontario L5R 4J3, Canada<br />

Wiley also publishes its books <strong>in</strong> a variety of electronic formats. Some content that appears <strong>in</strong> pr<strong>in</strong>t may not be available <strong>in</strong> electronic books.<br />

Library of Congress Catalog<strong>in</strong>g-<strong>in</strong>-Publication Data<br />

Darby, Stephen.<br />

<strong>River</strong> restoration : manag<strong>in</strong>g <strong>the</strong> uncerta<strong>in</strong>ty <strong>in</strong> restor<strong>in</strong>g physical habitat / Stephen Darby and David Sear.<br />

p. cm.<br />

Includes bibliographical references and <strong>in</strong>dex.<br />

ISBN 978-0-470-86706-8 (cloth)<br />

1. Stream restoration. 2. Riparian restoration. I. Sear, David (David A.) II. Title.<br />

TC409.D28 2008<br />

627′.12–dc22<br />

2007030210<br />

British Library Catalogu<strong>in</strong>g <strong>in</strong> Publication Data<br />

A catalogue record for this book is available from <strong>the</strong> British Library<br />

ISBN-13 978-0-470-86706-8 (HB)<br />

Typeset <strong>in</strong> 9/11 pt Times by SNP Best-set Typesetter Ltd., Hong Kong<br />

Pr<strong>in</strong>ted and bound <strong>in</strong> Great Brita<strong>in</strong> by Antony Rowe Ltd, Chippenham, Wiltshire


Contents<br />

Preface vii<br />

List of Contributors xi<br />

Section I Introduction: The Nature and Signifi cance of <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong><br />

<strong>River</strong> <strong>Restoration</strong><br />

1 <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 3<br />

J. Lemons and R. Victor<br />

2 Sources of <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> Research 15<br />

W. L. Graf<br />

3 The Scope of Uncerta<strong>in</strong>ties <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 21<br />

J.M. Wheaton, S.E. Darby and D.A. Sear<br />

Section II Plann<strong>in</strong>g and Design<strong>in</strong>g <strong>Restoration</strong> Projects<br />

4 Plann<strong>in</strong>g <strong>River</strong> <strong>Restoration</strong> Projects: Social and Cultural Dimensions 43<br />

G.M. Kondolf and C-N. Yang<br />

5 Conceptual and Ma<strong>the</strong>matical Modell<strong>in</strong>g <strong>in</strong> <strong>River</strong> <strong>Restoration</strong>: Do We Have<br />

Unreasonable Confi dence? 61<br />

M. Stewardson and I. Ru<strong>the</strong>rfurd<br />

6 <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 79<br />

F.M.R. Hughes, T. Moss and K.S. Richards<br />

7 Hydrological and Hydraulic Aspects of <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 105<br />

N.J. Clifford, M.C. Acreman and D.J. Booker<br />

8 <strong>Uncerta<strong>in</strong>ty</strong> Surround<strong>in</strong>g <strong>the</strong> Ecological Targets and Response of <strong>River</strong> and Stream <strong>Restoration</strong> 139<br />

M.R. Perrow, E.R. Skeate, D. Leem<strong>in</strong>g, J. England and M.L. Toml<strong>in</strong>son<br />

Section III The Construction and Post-Construction Phases<br />

9 Construct<strong>in</strong>g <strong>Restoration</strong> Schemes: <strong>Uncerta<strong>in</strong>ty</strong>, Challenges and Opportunities 167<br />

J. Mant, R. Richardson and M. Janes


vi Contents<br />

10 Measures of Success: <strong>Uncerta<strong>in</strong>ty</strong> and Defi n<strong>in</strong>g <strong>the</strong> Outcomes of <strong>River</strong> <strong>Restoration</strong> Schemes 187<br />

K. Sk<strong>in</strong>ner, F.D. Shields, Jr and S. Harrison<br />

11 Methods for Evaluat<strong>in</strong>g <strong>the</strong> Geomorphological Performance of Naturalized <strong>River</strong>s:<br />

Examples from <strong>the</strong> Chicago Metropolitan Area 209<br />

B.L. Rhoads, M.H. Garcia, J. Rodriguez, F. Bombardelli, J. Abad and M. Daniels<br />

12 <strong>Uncerta<strong>in</strong>ty</strong> and <strong>the</strong> Management of <strong>Restoration</strong> Projects: The Construction and Early<br />

Post-Construction Phases 229<br />

A. Brookes and H. Dangerfi eld<br />

Section IV <strong>Uncerta<strong>in</strong>ty</strong> and Susta<strong>in</strong>ability: <strong>Restoration</strong> <strong>in</strong> <strong>the</strong> Long Term<br />

13 The Susta<strong>in</strong>ability of Restored <strong>River</strong>s: Catchment-Scale Perspectives on Long Term Response 253<br />

K.J. Gregory and P.W. Downs<br />

14 <strong>Uncerta<strong>in</strong>ty</strong> and The Susta<strong>in</strong>able Management of Restored <strong>River</strong>s 287<br />

M.D. Newson and M.J. Clark<br />

Index 303


For many years scientists and river practitioners have<br />

recognised <strong>the</strong> severity and extent to which aquatic ecosystems<br />

have been degraded by a variety of human disturbances<br />

and activities (Gregory and Park, 1974; Sear and<br />

Arnell, 2006). In turn, realisation of <strong>the</strong> widespread nature<br />

of <strong>the</strong> problem has more recently elicited a surge of <strong>in</strong>terest<br />

<strong>in</strong> <strong>the</strong> possibility of undertak<strong>in</strong>g corrective <strong>in</strong>terventions,<br />

such as fl ow restoration and channel modifi cations,<br />

to restore or rehabilitate lost and/or damaged ecosystem<br />

functions (Brookes and Shields, 1996; Wissmar and<br />

Bisson, 2003). Indeed, <strong>the</strong>re is now a substantial volume<br />

of literature on <strong>the</strong> broad topic of river restoration, much<br />

of which suggests that, to be susta<strong>in</strong>able, river restoration<br />

projects should be designed to recreate functional characteristics<br />

with<strong>in</strong> a context of physical (i.e. geomorphic)<br />

stability. It is true that <strong>the</strong> emphasis on stable channel<br />

design may refl ect <strong>the</strong> traditional discipl<strong>in</strong>es of many of<br />

<strong>the</strong> river eng<strong>in</strong>eers who have now turned <strong>the</strong>ir attentions<br />

to restoration. Whatever <strong>the</strong> provenance and merits of this<br />

approach, a focus on stable channel design requires <strong>the</strong><br />

application of geomorphic and eng<strong>in</strong>eer<strong>in</strong>g design tools<br />

(models) that are for <strong>the</strong> most part ei<strong>the</strong>r entirely empirical<br />

or empirically calibrated. As a result, different results are<br />

obta<strong>in</strong>ed when different models are applied to <strong>the</strong> same<br />

problem. Fur<strong>the</strong>rmore, <strong>the</strong> data required to apply morphological<br />

models to restoration design are often absent,<br />

<strong>in</strong>complete, or subject to measurement error. F<strong>in</strong>ally, even<br />

when a restoration design is completed, it is usually not<br />

possible to predict <strong>the</strong> precise sequence of fl ood events.<br />

In <strong>the</strong> long term fur<strong>the</strong>r variability is <strong>in</strong>troduced by climatic<br />

or catchment changes (e.g. <strong>in</strong> land use), or unanticipated<br />

social or cultural changes, all of which might shift<br />

<strong>the</strong> basic premise(s) of <strong>the</strong> design. It is evident that <strong>the</strong><br />

designers and managers of stream restoration projects are<br />

<strong>in</strong>evitably confronted with uncerta<strong>in</strong>ty.<br />

Despite, or perhaps because of, this challeng<strong>in</strong>g situation<br />

<strong>the</strong> restoration literature, albeit with some notable<br />

Preface<br />

exceptions (Wissmar and Bisson, 2003), has not yet<br />

devoted consideration to identify<strong>in</strong>g associated uncerta<strong>in</strong>ties,<br />

let alone seek<strong>in</strong>g to quantify, manage, or – where<br />

appropriate (see below) – constra<strong>in</strong> <strong>the</strong>m. Ra<strong>the</strong>r, <strong>the</strong> discipl<strong>in</strong>e<br />

has <strong>in</strong>stead tended to focus on management<br />

responses (e.g. post-project appraisal, adaptive management<br />

strategies) that only implicitly confront assumed<br />

sources of variability and uncerta<strong>in</strong>ty. Our concern is that<br />

a collective discipl<strong>in</strong>ary failure to recognise, communicate<br />

and deal appropriately with uncerta<strong>in</strong>ties might, at some<br />

time <strong>in</strong> <strong>the</strong> future, underm<strong>in</strong>e <strong>in</strong>stitutional and public confi<br />

dence <strong>in</strong> river restoration. In a fi rst attempt to address<br />

<strong>the</strong>se issues, we (toge<strong>the</strong>r with Dr Andrew Collison and<br />

Dr Sean Bennett) convened a special session on <strong>Uncerta<strong>in</strong>ty</strong><br />

<strong>in</strong> <strong>River</strong> <strong>Restoration</strong> at <strong>the</strong> 2002 Fall Meet<strong>in</strong>g of <strong>the</strong><br />

American Geophysical Union (AGU) <strong>in</strong> San Francisco,<br />

California. While recognis<strong>in</strong>g that no s<strong>in</strong>gle volume can<br />

ever cover all aspects of such a multi-faceted discipl<strong>in</strong>e as<br />

river restoration, <strong>the</strong> positive response to <strong>the</strong> topic at that<br />

AGU symposium prompted us to seek to explore it fur<strong>the</strong>r<br />

<strong>in</strong> this volume. All <strong>the</strong> chapters for this book were, <strong>the</strong>refore,<br />

specially commissioned <strong>in</strong> an attempt to provide a<br />

coherent narrative structure that offers a rational <strong>the</strong>oretical<br />

analysis of <strong>the</strong> uncerta<strong>in</strong> basis of restoration, while<br />

simultaneously provid<strong>in</strong>g practical guidance on manag<strong>in</strong>g<br />

<strong>the</strong> implications of that uncerta<strong>in</strong>ty.<br />

The result<strong>in</strong>g book is structured <strong>in</strong>to four ma<strong>in</strong> sections.<br />

Each offers a range of case studies <strong>in</strong> an attempt to ensure<br />

a wide geographic coverage. Likewise, <strong>the</strong> authorship is<br />

drawn from a range of countries and discipl<strong>in</strong>es, <strong>in</strong> an<br />

attempt to br<strong>in</strong>g a range of perspectives to <strong>the</strong> table.<br />

Section I comprises three chapters that review <strong>the</strong> nature<br />

and signifi cance of uncerta<strong>in</strong>ty <strong>in</strong> river restoration, provid<strong>in</strong>g<br />

a context for <strong>the</strong> rema<strong>in</strong>der of <strong>the</strong> book. In Chapter 1<br />

Lemons and Victor focus on <strong>the</strong> specifi c nature of scientifi<br />

c uncerta<strong>in</strong>ty <strong>in</strong> restoration, while Graf (Chapter 2)<br />

expands on this <strong>the</strong>me, identify<strong>in</strong>g a series of sources of


viii Preface<br />

uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong>ory, research and communication. In<br />

Chapter 3 Wheaton et al. syn<strong>the</strong>sise and extend <strong>the</strong>se<br />

analyses, present<strong>in</strong>g a classifi cation that suggests uncerta<strong>in</strong>ty<br />

fundamentally arises ei<strong>the</strong>r through limited knowledge<br />

or through natural system variability. This is an<br />

important dist<strong>in</strong>ction, not least because it helps to discrim<strong>in</strong>ate<br />

between those sources of uncerta<strong>in</strong>ty (limited<br />

knowledge) which should, where possible, be constra<strong>in</strong>ed<br />

(e.g. by scientifi c progress) from those sources (e.g. natural<br />

variability) that should be embraced to promote healthy<br />

system function<strong>in</strong>g. The management implications associated<br />

with each form of uncerta<strong>in</strong>ty are <strong>the</strong>refore dist<strong>in</strong>ct,<br />

but recognition that embrac<strong>in</strong>g certa<strong>in</strong> types of uncerta<strong>in</strong>ty<br />

may be both necessary and desirable to assure susta<strong>in</strong>ability<br />

is a <strong>the</strong>me that runs throughout many of <strong>the</strong> contributions<br />

here<strong>in</strong>.<br />

The book is subsequently structured to address <strong>the</strong> discrete<br />

stages <strong>in</strong> <strong>the</strong> life span of a typical restoration project,<br />

cover<strong>in</strong>g <strong>the</strong> plann<strong>in</strong>g and design activities associated with<br />

<strong>the</strong> pre-construction phase (Section II), <strong>the</strong> construction<br />

phase itself (Section III) and <strong>the</strong> long term post-construction<br />

phase (Section IV). Section II (Chapters 4 to 8) presents<br />

contributions cover<strong>in</strong>g various aspects of plann<strong>in</strong>g<br />

and design associated with restoration projects. In Chapter<br />

4 Kondolf and Yang’s review rem<strong>in</strong>ds us that restoration<br />

is fundamentally a social and cultural process, with variability<br />

<strong>in</strong> cultural values act<strong>in</strong>g as a signifi cant contributor<br />

of uncerta<strong>in</strong>ty. Present<strong>in</strong>g an Australian case study, where<br />

<strong>the</strong> aim was to restore fl ows capable of fl ush<strong>in</strong>g fi ne sediment<br />

from river gravels, Stewardson et al. (Chapter 5)<br />

identify <strong>the</strong> limits <strong>in</strong> our understand<strong>in</strong>g of hydrological,<br />

hydraulic and geomorphic processes and how <strong>the</strong>se constra<strong>in</strong><br />

our ability to model river system dynamics. In terms<br />

of <strong>the</strong> uncerta<strong>in</strong>ty classifi cation discussed <strong>in</strong> Chapter 3,<br />

<strong>the</strong>ir focus is essentially on quantify<strong>in</strong>g <strong>the</strong> magnitude of<br />

uncerta<strong>in</strong>ty due to limited knowledge. Their results (that<br />

<strong>the</strong> magnitude of designed fl ush<strong>in</strong>g fl ows is subject to<br />

uncerta<strong>in</strong>ty estimates approximately twice that of <strong>the</strong> fl ow<br />

itself) re<strong>in</strong>force <strong>the</strong> earlier suggestion that restoration is<br />

<strong>in</strong>deed an uncerta<strong>in</strong> discipl<strong>in</strong>e. Whe<strong>the</strong>r this really means<br />

that we should have ‘unreasonable confi dence’ <strong>in</strong> restoration,<br />

as suggested by <strong>the</strong>ir provocative sub-title, is a <strong>the</strong>me<br />

that is cont<strong>in</strong>ued throughout <strong>the</strong> book.<br />

In contrast to <strong>the</strong> focus on uncerta<strong>in</strong>ty due to limited<br />

knowledge expounded <strong>in</strong> Chapter 5, Chapter 6 (Hughes<br />

et al.) reviews some of <strong>the</strong> diffi culties associated with<br />

extend<strong>in</strong>g restoration <strong>in</strong>to complex riparian and fl oodpla<strong>in</strong><br />

habitats, emphasis<strong>in</strong>g that <strong>in</strong> <strong>the</strong>se systems uncerta<strong>in</strong>ty (<strong>in</strong><br />

this case <strong>in</strong> <strong>the</strong> form of physical diversity and variability)<br />

is necessary to underp<strong>in</strong> <strong>the</strong> successful restoration of<br />

forest fl oodpla<strong>in</strong> ecosystems. In Chapter 7, Clifford et al.’s<br />

comprehensive review of how <strong>the</strong> restoration of fl ow<br />

hydrology and hydraulics can be used to enhance aquatic<br />

habitats also recognises <strong>the</strong> importance of restor<strong>in</strong>g natural<br />

variability, and provides recommendations on how such<br />

variability can be <strong>in</strong>terpreted <strong>in</strong> modell<strong>in</strong>g <strong>in</strong>vestigations.<br />

The <strong>the</strong>me of uncerta<strong>in</strong>ty associated with ecological<br />

targets is explored fur<strong>the</strong>r <strong>in</strong> Chapter 8 (Perrow et al.),<br />

where <strong>the</strong> paradox that uncerta<strong>in</strong>ty is often viewed as a<br />

pejorative term is aga<strong>in</strong> highlighted, even if it is uncerta<strong>in</strong>ty<br />

(<strong>in</strong> <strong>the</strong> form of natural variability) that is <strong>the</strong> key<br />

mechanism for susta<strong>in</strong><strong>in</strong>g healthy ecosystems. How uncerta<strong>in</strong>ty<br />

due to <strong>the</strong> lack of understand<strong>in</strong>g of a discipl<strong>in</strong>e<br />

(ecology) by river managers has led to a lack of us<strong>in</strong>g<br />

available science with<strong>in</strong> restoration process is also<br />

highlighted.<br />

Section III (Chapters 9 to 12) addresses <strong>the</strong> construction<br />

phase of a restoration project, which is defi ned <strong>in</strong> this book<br />

as extend<strong>in</strong>g up to one or two years after completion of<br />

<strong>the</strong> project. The contributions <strong>in</strong> this section are written<br />

primarily by river practitioners, who employ <strong>the</strong>ir collective<br />

experience to offer a range of perspectives on uncerta<strong>in</strong>ties<br />

encountered dur<strong>in</strong>g this key stage of restoration.<br />

Mant et al. (Chapter 9) review <strong>the</strong> diffi culties encountered<br />

dur<strong>in</strong>g construction and note that strong teamwork skills<br />

are required to ensure that <strong>the</strong> design concepts provided<br />

by geomorphologists and ecologists are correctly translated<br />

<strong>in</strong>to practice by those responsible for construction.<br />

A diffi culty here is that restoration is seen as a relatively<br />

new facet of civil eng<strong>in</strong>eer<strong>in</strong>g, such that contractors may<br />

not always have <strong>the</strong> experience necessary to recognise that<br />

variability, ra<strong>the</strong>r than uniformity (<strong>the</strong>ir experience to<br />

date), is often necessary. To this end it is essential that<br />

designers <strong>in</strong>form <strong>the</strong> workforce of <strong>the</strong> specifi c requirements<br />

of <strong>the</strong> river restoration project, while project managers<br />

must also take responsibility for monitor<strong>in</strong>g<br />

construction as it progresses. This po<strong>in</strong>ts to <strong>the</strong> importance<br />

of ensur<strong>in</strong>g that <strong>the</strong> constructed project does <strong>in</strong>deed<br />

conform to <strong>the</strong> design specifi cations, rais<strong>in</strong>g <strong>the</strong> issue of<br />

evaluat<strong>in</strong>g project outcomes.<br />

This subject is <strong>the</strong> <strong>the</strong>me of <strong>the</strong> next three chapters.<br />

Sk<strong>in</strong>ner et al. (Chapter 10) review post-project appraisals<br />

with reference to both physical and ecological measures<br />

of success, whilst Rhoads et al. (Chapter 11) focus on<br />

methods for evaluat<strong>in</strong>g <strong>the</strong> geomorphological performance<br />

of restored rivers, provid<strong>in</strong>g examples from heavily urbanised<br />

catchments <strong>in</strong> Ill<strong>in</strong>ois <strong>in</strong> <strong>the</strong> USA. Both contributions<br />

emphasise <strong>the</strong> key need for both pre and post-project monitor<strong>in</strong>g,<br />

even if only to a m<strong>in</strong>imum standard. This is viewed<br />

as necessary to verify that projects are constructed accord<strong>in</strong>g<br />

to <strong>the</strong>ir design, as well as an <strong>in</strong>tegral tool of adaptive<br />

management that can make project adjustments <strong>in</strong> <strong>the</strong> face<br />

of uncerta<strong>in</strong>ties <strong>in</strong>troduced by variable post-project conditions.<br />

This <strong>the</strong>me is fur<strong>the</strong>r explored by Brookes and


Dangerfi eld (Chapter 12), who propose that managers<br />

should adopt cont<strong>in</strong>uous improvement as an overall operational<br />

philosophy for restor<strong>in</strong>g rivers. The term cont<strong>in</strong>uous<br />

improvement is widely documented <strong>in</strong> human resource,<br />

organisational management and environmental management<br />

literature, and is taken to be a philosophy of learn<strong>in</strong>g<br />

dur<strong>in</strong>g <strong>the</strong> construction and post-construction phases (and<br />

mak<strong>in</strong>g adaptations to a particular project as necessary)<br />

for <strong>the</strong> benefi t of <strong>the</strong> cont<strong>in</strong>u<strong>in</strong>g work and future practice.<br />

This appears to be a robust approach that has <strong>the</strong> potential,<br />

over time, to reduce uncerta<strong>in</strong>ties associated with limited<br />

knowledge while simultaneously provid<strong>in</strong>g a framework<br />

for adaptive response to uncerta<strong>in</strong>ties associated with<br />

natural variability.<br />

Section IV (Chapters 13 and 14) addresses <strong>the</strong> challenge<br />

of <strong>the</strong> need to assure <strong>the</strong> long term susta<strong>in</strong>ability of restoration<br />

projects <strong>in</strong> <strong>the</strong> face of uncerta<strong>in</strong> futures. There is a<br />

clear recognition that as <strong>the</strong> time scales over which project<br />

outcomes should be considered <strong>in</strong>crease, <strong>the</strong>re is a concomitant<br />

need to address <strong>in</strong>creased spatial scales. Specifi -<br />

cally, <strong>the</strong>re is a need to consider how catchment-scale<br />

processes (which <strong>in</strong>fl uence <strong>the</strong> fl uxes of water and sediment<br />

supplied to restoration reaches) are to be susta<strong>in</strong>ed<br />

<strong>in</strong> <strong>the</strong> long term. Clearly, as spatial and temporal scales<br />

<strong>in</strong>crease, <strong>the</strong>n so do <strong>the</strong> uncerta<strong>in</strong>ties particularly, but not<br />

exclusively so, those associated with <strong>in</strong>creases <strong>in</strong> spatial<br />

and temporal variability. In Chapter 13 Downs and Gregory<br />

br<strong>in</strong>g a hydromorphological perspective to <strong>the</strong>se issues,<br />

suggest<strong>in</strong>g that <strong>the</strong> bounds of <strong>the</strong>se uncerta<strong>in</strong>ties can be<br />

evaluated with reference to long term (palaeo)hydrological<br />

and geomorphological evidence of past catchment response<br />

– <strong>in</strong> effect advocat<strong>in</strong>g a more precise defi nition of <strong>the</strong><br />

uncerta<strong>in</strong>ty due to natural variability.<br />

The fi nal chapter (Chapter 14; Newson & Clark) provides<br />

an apt conclusion. Recognis<strong>in</strong>g that uncerta<strong>in</strong>ty (<strong>in</strong><br />

<strong>the</strong> form of natural variability) is both endemic and necessary,<br />

it is noted that <strong>the</strong>re is a confl ict between <strong>the</strong> precautionary<br />

pr<strong>in</strong>ciple – a cornerstone of susta<strong>in</strong>able th<strong>in</strong>k<strong>in</strong>g<br />

– and uncerta<strong>in</strong>ty. Newson and Clark recognise that all<br />

restorations have outcomes that are to some extent unpredictable,<br />

and <strong>the</strong> precautionary pr<strong>in</strong>ciple thus becomes a<br />

recipe for <strong>in</strong>action. <strong>Uncerta<strong>in</strong>ty</strong> is <strong>the</strong>refore simultaneously<br />

necessary for, but also a barrier to, susta<strong>in</strong>ability.<br />

They attempt to resolve this particular problem by identify<strong>in</strong>g<br />

management and restoration opportunities that are<br />

susta<strong>in</strong>able despite be<strong>in</strong>g uncerta<strong>in</strong>, not<strong>in</strong>g that <strong>in</strong> practical<br />

terms it is to adaptive management that we most often turn<br />

for a way forward, re<strong>in</strong>forc<strong>in</strong>g a series of conclusions from<br />

earlier chapters.<br />

Do we have unreasonable confi dence <strong>in</strong> restoration,<br />

based on <strong>the</strong> state of <strong>the</strong> art, or are we happy to boldly go<br />

Preface ix<br />

with <strong>the</strong> uncerta<strong>in</strong> ebb and fl ow of natural variability?<br />

Perhaps <strong>the</strong> way forward is through a clearer and more<br />

transparent approach to communicat<strong>in</strong>g uncerta<strong>in</strong>ty, such<br />

that all participants – and especially stakeholders – understand<br />

that while every effort can and should be made to<br />

identify, and where possible reduce, scientifi c and methodological<br />

uncerta<strong>in</strong>ties, uncerta<strong>in</strong>ty due to natural variability<br />

is both welcome and necessary to susta<strong>in</strong> healthy<br />

aquatic ecosystems. This implies a concerted approach on<br />

two fronts: scientists can cont<strong>in</strong>ue to refi ne <strong>the</strong> knowledge<br />

base (and managers need to recognise <strong>the</strong> value of this<br />

research for <strong>the</strong>ir (adaptive) management practices), while<br />

managers must work harder to build and accommodate<br />

variability <strong>in</strong>to projects. <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> river restoration is<br />

endemic but clearly offers opportunities, not just as a<br />

rationale for fur<strong>the</strong>r research, but fundamentally for more<br />

susta<strong>in</strong>ably managed restoration projects. Just as life goes<br />

on with little confi dence or ability to predict <strong>the</strong> future, so<br />

restoration must cont<strong>in</strong>ue to evolve and adopt an approach<br />

that is consistent with <strong>the</strong> uncerta<strong>in</strong> function<strong>in</strong>g of river<strong>in</strong>e<br />

ecosystems.<br />

In clos<strong>in</strong>g this preface, we would like to acknowledge<br />

those who have made signifi cant contributions dur<strong>in</strong>g <strong>the</strong><br />

production of this book. Firstly, we would like to thank<br />

those numerous professionals who provided detailed peer<br />

reviews of each chapter, often work<strong>in</strong>g to tight deadl<strong>in</strong>es.<br />

The anonymous nature of peer review means that we are<br />

unable to identify <strong>the</strong>m here, but you know who you are!<br />

F<strong>in</strong>ally, Tim Aspden and <strong>the</strong> staff of <strong>the</strong> Cartographic Unit<br />

at <strong>the</strong> School of Geography, University of Southampton,<br />

provided guidance on, and help with, <strong>the</strong> production of<br />

much of <strong>the</strong> artwork. A fi nal acknowledgment must go to<br />

our families who have put up with longer hours than<br />

normal (or natural!) <strong>in</strong> <strong>the</strong> drive for completion.<br />

Stephen Darby and David Sear<br />

December 2007<br />

REFERENCES<br />

Brookes A, Shields FD. 1996. <strong>River</strong> Channel <strong>Restoration</strong>:<br />

Guid<strong>in</strong>g Pr<strong>in</strong>ciples for Susta<strong>in</strong>able Projects. John Wiley &<br />

Sons Ltd: Chichester, UK.<br />

Gregory KJ, Park CC. 1974. Adjustment of river channel capacity<br />

downstream from a reservoir. Water Resources Research 10:<br />

840–873.<br />

Sear DA, Arnell NW. 2006. The application of palaeohydrology<br />

to river management. Catena 66: 169–183.<br />

Wissmar RC, Bisson PA (Eds). 2003c. Strategies for Restor<strong>in</strong>g<br />

<strong>River</strong> Ecosystems: Sources of Variability and <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong><br />

Natural and Managed Systems. American Fisheries Society:<br />

Be<strong>the</strong>sda, Maryland, USA.


Jorge Abad, Department of Civil and Environmental Eng<strong>in</strong>eer<strong>in</strong>g,<br />

University of Ill<strong>in</strong>ois at Urbana-Champaign,<br />

Urbana, Ill<strong>in</strong>ois 61801, USA.<br />

Dr Mike Acreman, Water Resources & Environment Division,<br />

Centre for Ecology & Hydrology, Maclean Build<strong>in</strong>g,<br />

Crowmarsh Gifford, Wall<strong>in</strong>gford, Oxfordshire OX10 8BB,<br />

UK (man@ceh.ac.uk).<br />

Professor Fabian Bombadelli, Department of Civil and<br />

Environmental Eng<strong>in</strong>eer<strong>in</strong>g, University of California at<br />

Davis, Davis, California 95616, USA (fabombardelli@<br />

ucdavis.edu).<br />

Dr Douglas Booker, National Institute of Water and<br />

Atmospheric Research, 10 Kyle St., Riccarton, Christchurch,<br />

8011, New Zealand (d.booker@niwa.co.nz).<br />

Dr Andrew Brookes, Jacobs (UK), School Green, Sh<strong>in</strong>fi<br />

eld, Read<strong>in</strong>g RG2 9HL, UK (andrew.brookes@jacobs.<br />

com).<br />

Professor Mike Clark, School of Geography, University of<br />

Southampton, Highfi eld, Southampton SO17 1BJ, UK<br />

(M.J.Clark@soton.ac.uk).<br />

Professor Nick Clifford, School of Geography, The University<br />

of Nott<strong>in</strong>gham, University Park, Nott<strong>in</strong>gham NG7<br />

2RD, UK (Nick.Clifford@nott<strong>in</strong>gham.ac.uk).<br />

Dr Helen Dangerfi eld, Royal Haskon<strong>in</strong>g, 4 Dean’s<br />

Yard, London, SW19 3NL, UK (Helen.dangerfi eld@<br />

royalhaskon<strong>in</strong>g.com).<br />

List of Contributors<br />

Dr Mel<strong>in</strong>da Daniels, Department of Geography, Kansas<br />

State University, USA (mel<strong>in</strong>da.daniels@uconn.edu).<br />

Dr Stephen Darby, School of Geography, University of<br />

Southampton, Highfi eld, Southampton SO17 1BJ, UK<br />

(S.E.Darby@soton.ac.uk).<br />

Dr Peter Downs, Stillwater Sciences, 2855 Telegraph<br />

Avenue, #400, Berkeley, California 94705 USA (downs@<br />

stillwatersci.com).<br />

Dr Judy England, Environment Agency, Apollo Court, 2<br />

Bishop’s Square, St. Albans Road West, Hatfi eld AL10<br />

9EX, UK.<br />

Professor Marcelo Garcia, Department of Civil and Environmental<br />

Eng<strong>in</strong>eer<strong>in</strong>g, University of Ill<strong>in</strong>ois at Urbana-<br />

Champaign, Urbana, Ill<strong>in</strong>ois 61801, USA (mhgarcia@<br />

uiuc.edu).<br />

Professor William Graf, Department of Geography, University<br />

of South Carol<strong>in</strong>a, Columbia, South Carol<strong>in</strong>a<br />

29208, USA (graf@sc.edu).<br />

Professor Ken Gregory, School of Geography, University<br />

of Southampton, Highfi eld, Southampton SO17 1BJ, UK<br />

(Ken.Gregory@bt<strong>in</strong>ternet.com).<br />

Dr Simon Harrison, Department of Zoology, Ecology and<br />

Plant Sciences, University College Cork, Lee Malt<strong>in</strong>gs,<br />

Prospect Row, Cork, Ireland (s.harrison@ucc.ie).


xii List of Contributors<br />

Dr Franc<strong>in</strong>e Hughes, Department of Life Sciences, Anglia<br />

Rusk<strong>in</strong> University, East Road, Cambridge CB1 1PT, UK<br />

(f.hughes@anglia.ac.uk).<br />

Mart<strong>in</strong> Janes, <strong>River</strong> <strong>Restoration</strong> Centre Manager, The<br />

<strong>River</strong> <strong>Restoration</strong> Centre, Build<strong>in</strong>g 53, Cranfi eld University,<br />

Cranfi eld, Bedfordshire MK43 0AL, UK (rrc@<strong>the</strong>rrc.<br />

co.uk).<br />

Professor Mat Kondolf, Department of Landscape Architecture<br />

and Environmental Plann<strong>in</strong>g, University of California,<br />

Berkeley, California 94720-2000, USA (kondolf@<br />

ucl<strong>in</strong>k.berkeley.edu).<br />

David Leem<strong>in</strong>g, Consultant Ecologist, Sp<strong>in</strong>dlewood, 45<br />

West End, Ashwell, Hertfordshire SG7 5QY, UK.<br />

Professor John Lemons, Department of Environmental<br />

Studies, University of New England,11 Hills Beach Road,<br />

Biddeford, Ma<strong>in</strong>e 04005, USA (jlemons@une.edu).<br />

Dr Jenny Mant, The <strong>River</strong> <strong>Restoration</strong> Centre, Build<strong>in</strong>g<br />

53, Cranfi eld University, Cranfi eld, Bedfordshire MK43<br />

0AL, UK (rrc@<strong>the</strong>rrc.co.uk).<br />

Dr Tim Moss, Institute for Regional Development and<br />

Structural Plann<strong>in</strong>g (IRS), Flakenstrasse 28–31, 15537<br />

Erkner, Germany. (MossT@irs-net.de).<br />

Professor Malcolm Newson, Department of Geography,<br />

University of Newcastle-Upon-Tyne, Newcastle-Upon-<br />

Tyne NE1 7RU, UK (M.D.Newson@ncl.ac.uk).<br />

Dr Mart<strong>in</strong> Perrow, ECON, Ecological Consultancy,<br />

Norwich Research Park, Colney Lane, Norwich, Norfolk<br />

NR4 7UH, UK (m.perrow@econ-ecology.com).<br />

Professor Bruce Rhoads, Department of Geography, University<br />

of Ill<strong>in</strong>ois at Urbana-Champaign, Room 220 Davenport<br />

Hall, 607 South Ma<strong>the</strong>ws Avenue, Urbana, Ill<strong>in</strong>ois<br />

61801-3671, USA (brhoads@uiuc.edu).<br />

Professor Keith Richards, Department of Geography, University<br />

of Cambridge, Down<strong>in</strong>g Place, Cambridge CB2<br />

3EN, UK (keith.richards@geog.cam.ac.uk).<br />

Dr Roy Richardson, Scottish Environment Protection<br />

Agency, Burnbrae, Mossilee Road, Galashiels TD1 1NF,<br />

UK.<br />

Dr Jose Rodriguez, Faculty of Eng<strong>in</strong>eer<strong>in</strong>g and Built Environment,<br />

University of Newcastle, Newcastle, New South<br />

Wales 2308, Australia (jose.rodriguez@newcastle.edu.<br />

au).<br />

Dr Ian Ru<strong>the</strong>rfurd, Geography Program, School of<br />

Resource Management, University of Melbourne, Melbourne,<br />

Victoria 3010 Australia (idruth@unimelb.edu.au).<br />

Professor David Sear, School of Geography, University of<br />

Southampton, Highfi eld, Southampton SO17 1BJ, UK<br />

(D.Sear@soton.ac.uk).<br />

Dr F. Doug Shields Jr, Water Quality and Ecology<br />

Research Unit, National Sedimentation Laboratory, USDA<br />

Agricultural Research Service, National Sedimentation<br />

Laboratory, PO Box 1157, Oxford, Mississippi, USA<br />

(dshields@ars.usda.gov).<br />

Eleanor R. Skeate, ECON, Ecological Consultancy,<br />

Norwich Research Park, Colney Lane, Norwich, Norfolk<br />

NR4 7UH, UK (e.skeate@econ-ecology.com).<br />

Dr Kev<strong>in</strong> Sk<strong>in</strong>ner, Pr<strong>in</strong>cipal Geomorphologist, Jacobs<br />

Babtie, School Green, Sh<strong>in</strong>fi eld, Read<strong>in</strong>g RG2 9HL UK<br />

(Kev<strong>in</strong>.sk<strong>in</strong>ner@jacobs.com).<br />

Dr Michael Stewardson, Department of Civil and Environmental<br />

Eng<strong>in</strong>eer<strong>in</strong>g and eWater CRC, University of<br />

Melbourne, Melbourne, Victoria, 3010 Australia (mjstew@<br />

unimelb.edu.au).<br />

Mark L. Toml<strong>in</strong>son ECON, Ecological Consultancy,<br />

Norwich Research Park, Colney Lane, Norwich, Norfolk<br />

NR4 7UH, UK (m.toml<strong>in</strong>son@econ-ecology.com).<br />

Professor Reg<strong>in</strong>ald Victor, Centre for Environmental<br />

Studies and Research, c/o Department of Biology, Sultan<br />

Qaboos University, PO Box 36, Al-Khod, PC 123, Muscat<br />

Sultanate of Oman (rvictor@squ.edu.om).<br />

Joseph M. Wheaton, Institute of Geography and Earth<br />

Sciences, University of Wales, Aberystwyth SY23 3DB,<br />

UK (Joe.Wheaton@aber.ac.uk).<br />

Dr Chia-N<strong>in</strong>g Yang, Department of Landscape Ar -<br />

chitecture, California State Polytechnic University,<br />

Pomona, California 91768, USA (cnyang@csupomona.<br />

edu).


SECTION I<br />

Introduction: The Nature and Signifi cance<br />

of <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>River</strong> <strong>Restoration</strong>


<strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat Edited by Stephen Darby and David Sear<br />

© 2008 John Wiley & Sons, Ltd<br />

1<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>River</strong> <strong>Restoration</strong><br />

John Lemons 1 and Reg<strong>in</strong>ald Victor 2<br />

1 Department of Environmental Studies, University of New England, USA<br />

2 Center for Environmental Studies and Research, Department of Biology, Sultan Qaboos University, Sultanate of Oman<br />

1.1 INTRODUCTION<br />

As we are well aware, rivers fundamentally shape <strong>the</strong><br />

planet and human life. Both ancient and modern societies<br />

have developed and fl ourished <strong>in</strong> <strong>the</strong> proximity of rivers<br />

and this trend has cont<strong>in</strong>ued till modern times. Nienhuis<br />

and Leuven (2001) summarize how humans have spatially<br />

and temporally altered rivers over a 6000-year period by<br />

various anthropogenic activities. For example, <strong>in</strong>tensive<br />

use of European rivers started over 500 years ago lead<strong>in</strong>g<br />

to <strong>the</strong> loss of <strong>the</strong>ir ecological <strong>in</strong>tegrity (Smits et al., 2000).<br />

Some rivers were altered for navigation, fl ood control,<br />

agriculture and reclamation of land for urban development,<br />

while most were used as chutes for waste disposal<br />

<strong>in</strong>clud<strong>in</strong>g sewage, <strong>the</strong>rmal effl uents and both nontoxic and<br />

toxic chemicals; some rivers were also rout<strong>in</strong>ely dredged<br />

to facilitate <strong>the</strong> transport and storage of timber, while<br />

o<strong>the</strong>rs were heavily fi shed (Ward and Stanford, 1979; De<br />

Wall et al., 1995; Eiseltova and Biggs, 1995).<br />

Large river systems (stream order >8) all over <strong>the</strong> world<br />

have been extensively dammed for hydroelectric power,<br />

recreation, fl ood control and to divert water to support<br />

agriculture. Impacts of large dams <strong>in</strong>clude <strong>the</strong> loss of<br />

fi sheries and <strong>the</strong> ecological collapse of <strong>the</strong> entire river<br />

regime (Balon and Coche, 1974; Rzoska, 1976; Obeng,<br />

1981). Extensive series of levees built along large rivers<br />

have caused major losses of ecosystem structure and function.<br />

Along <strong>the</strong> Mississippi <strong>River</strong>, <strong>the</strong> largest river <strong>in</strong> North<br />

America, levees threaten federal plans to protect endangered<br />

species (EPA, 2004). The effects of impound<strong>in</strong>g<br />

small rivers (stream order 4–8) are even more drastic. In<br />

some West African small rivers entire fi sh communities<br />

had changed due to impoundment and <strong>the</strong> ecological perturbations<br />

extended for considerable distances downstream<br />

(Victor and Tetteh, 1988; Victor and Meye, 1994; Victor<br />

and Onomivbori, 1996). Gopal (2003) describes how<br />

rivers <strong>in</strong> arid and semi-arid regions <strong>in</strong> Asia are be<strong>in</strong>g<br />

degraded due to overexploitation of natural resources,<br />

sal<strong>in</strong>ization, pollution and <strong>in</strong>troduction of exotic species.<br />

Just as rivers have undergone alteration, so too have<br />

<strong>the</strong>re been efforts to restore <strong>the</strong>m <strong>in</strong> order to provide benefi<br />

ts to <strong>the</strong> environment and/or human health, as this book<br />

attests (see also MacMahon and Holl, 2001). Obviously,<br />

scientifi c research contributes to river restoration by: provid<strong>in</strong>g<br />

reliable and needed explanatory or heuristic knowledge<br />

and understand<strong>in</strong>g of restoration problems; help<strong>in</strong>g<br />

to identify and defi ne new research needs and directions<br />

through <strong>the</strong> acquisition of factual <strong>in</strong>formation; and <strong>in</strong>form<strong>in</strong>g<br />

policy and decision mak<strong>in</strong>g (Caldwell, 1996).<br />

A major premise of this book is that to be susta<strong>in</strong>able,<br />

river restoration projects need to effectively recreate a<br />

rivers’ functional characteristics tak<strong>in</strong>g <strong>in</strong>to account <strong>the</strong><br />

dynamic geomorphic characteristics. While many restoration<br />

projects have benefi ted environmental and/or human<br />

health, understudied sources of uncerta<strong>in</strong>ty limit confi -<br />

dence <strong>in</strong> predict<strong>in</strong>g <strong>the</strong> outcomes of restoration activities<br />

and programs. Specifi c examples of uncerta<strong>in</strong>ty <strong>in</strong> river<br />

restoration discussed <strong>in</strong> this book <strong>in</strong>clude those <strong>in</strong>herent<br />

<strong>in</strong>: river management processes; <strong>the</strong> plann<strong>in</strong>g and design<br />

phases of restoration projects; hydraulic and hydrological<br />

aspects of restoration; water quantity issues; identify<strong>in</strong>g<br />

appropriate ecological characteristics and predict<strong>in</strong>g <strong>the</strong>ir


4 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

responses <strong>in</strong> restoration designs; and <strong>the</strong> construction and<br />

post-construction phases of restoration projects. The<br />

sources of uncerta<strong>in</strong>ties <strong>in</strong>clude: lack of scientifi c and<br />

o<strong>the</strong>r <strong>in</strong>formation; limitations of analytical methods<br />

and tools; complexities of river systems; and needs<br />

to make value-laden judgments at all stages of river restoration<br />

problem identifi cation, analysis and solution<br />

implementation.<br />

Beg<strong>in</strong>n<strong>in</strong>g <strong>in</strong> and s<strong>in</strong>ce <strong>the</strong> early 1990s some philosophers,<br />

scientists and public policy experts concluded that<br />

<strong>the</strong> sources and implications of scientifi c and o<strong>the</strong>r uncerta<strong>in</strong>ty<br />

<strong>in</strong> environmental problem solv<strong>in</strong>g, <strong>in</strong>clud<strong>in</strong>g restoration,<br />

have been understudied and, as a consequence,<br />

not suffi ciently taken <strong>in</strong>to account by researchers, public<br />

policy makers and decision makers (Mayo and Hollander,<br />

1991; Cranor, 1993; Shrader-Frechette and McCoy, 1993;<br />

Funtowicz and Ravetz, 1995; Lemons and Brown, 1995;<br />

Lemons, 1996; EEA, 2001; Kriebel et al., 2001; Tickner,<br />

2002, 2003).<br />

In agree<strong>in</strong>g with this conclusion, <strong>the</strong> objective <strong>in</strong> this<br />

chapter is <strong>the</strong>refore to fi rst discuss various broad views<br />

about scientifi c uncerta<strong>in</strong>ty and <strong>in</strong>dicate how and why<br />

<strong>the</strong>se need to be taken <strong>in</strong>to greater account by scientists,<br />

policy makers and decision makers. (O<strong>the</strong>r chapters<br />

address uncerta<strong>in</strong>ty and analyze <strong>in</strong> more concrete detail<br />

how it <strong>in</strong>teracts with <strong>the</strong> specifi c <strong>the</strong>ories and practices of<br />

river restoration.). Discussion <strong>the</strong>n focuses on what might<br />

constitute ‘good’ science when science is used to <strong>in</strong>form<br />

policy and decision mak<strong>in</strong>g under conditions of scientifi c<br />

uncerta<strong>in</strong>ty. Value-laden sources and implications of<br />

uncerta<strong>in</strong>ty <strong>in</strong> river restoration are <strong>the</strong>n discussed because<br />

<strong>the</strong>y are both important but understudied. Discussion of<br />

<strong>the</strong> value-laden sources and implications of uncerta<strong>in</strong>ty is<br />

followed with: a brief discussion of some of <strong>the</strong> practical<br />

and policy implications of uncerta<strong>in</strong>ty <strong>in</strong> river restoration,<br />

and, fi nally, a brief case study of river restoration <strong>in</strong> order<br />

to communicate our views with a practical example. For<br />

reasons of brevity <strong>the</strong> case study communicates views<br />

about some, but not all, aspects of uncerta<strong>in</strong>ty <strong>in</strong> river<br />

restoration.<br />

Paren<strong>the</strong>tically, here it is necessary to comment on defi -<br />

nitions of ‘restoration’ when used <strong>in</strong> <strong>the</strong> context of river<br />

restoration. The fi eld of restoration ecology suffers from<br />

a lack of conceptual clarity concern<strong>in</strong>g its mean<strong>in</strong>g, goals<br />

and objectives. S<strong>in</strong>ce about <strong>the</strong> mid-1980s, <strong>the</strong> fi eld of<br />

river restoration has <strong>in</strong>creas<strong>in</strong>gly evolved <strong>in</strong> an attempt to<br />

better meet societies’ needs to more effectively repair<br />

damage to rivers (e.g., Cairns and Heckman, 1996; Karr<br />

and Chu, 1999; Cairns, 2001). The Society of Wetlands<br />

Scientists (SWS, 2000) defi ned restoration as ‘actions<br />

taken <strong>in</strong> a converted or degraded natural wetland that<br />

result <strong>in</strong> <strong>the</strong> re-establishment of ecological processes,<br />

function, and biotic/abiotic l<strong>in</strong>kages and lead to a persistent,<br />

resilient system <strong>in</strong>tegrated with<strong>in</strong> its landscape.’ In<br />

2002, <strong>the</strong> Society for Ecological <strong>Restoration</strong> (SER, 2002)<br />

defi ned restoration as <strong>the</strong> ‘. . . process of assist<strong>in</strong>g <strong>the</strong><br />

recovery of an ecosystem that has been degraded, damaged,<br />

or destroyed.’ Regardless of <strong>the</strong>se defi nitions, <strong>the</strong> goals<br />

and objectives of river restoration are not clear.<br />

Rolston (1988) believes that where possible ecosystems<br />

should be returned to <strong>the</strong>ir ‘natural’ or ‘orig<strong>in</strong>al’ condition.<br />

Westra (1995) argues that restoration should focus on<br />

restor<strong>in</strong>g ecosystems’ abilities to cont<strong>in</strong>ue <strong>the</strong>ir ongo<strong>in</strong>g<br />

change and development unconstra<strong>in</strong>ed by human <strong>in</strong>terruptions<br />

past or present. The United States National<br />

Research Council (NRC, 1999) defi ned restoration as<br />

‘<strong>the</strong> return of an ecosystem to a close approximation of<br />

its condition prior to disturbance.’ This defi nition was<br />

expanded by Cairns (2001), who asserted that <strong>the</strong> goal of<br />

restoration should be devoted to ‘return<strong>in</strong>g damaged ecosystems<br />

to a condition that is structurally and functionally<br />

similar to <strong>the</strong> predisturbance state.’<br />

Alternatively, o<strong>the</strong>rs <strong>in</strong>volved <strong>in</strong> <strong>the</strong> fi eld of restoration<br />

ecology provide defi nitions for restoration that more<br />

explicitly focus on historical, social, cultural, political,<br />

aes<strong>the</strong>tic and moral aspects. For example, Sweeney (2000)<br />

argues that restoration should focus on <strong>the</strong> value-laden<br />

social and ethical perspectives regard<strong>in</strong>g what constitutes<br />

a ‘restored’ ecosystem. Some o<strong>the</strong>rs ma<strong>in</strong>ta<strong>in</strong> that conservation<br />

and, by implication, restoration goals should take<br />

<strong>in</strong>to account <strong>the</strong> views and practices of rural and <strong>in</strong>digenous<br />

people who depend on <strong>the</strong> ecosystems for <strong>the</strong>ir<br />

physical and cultural subsistence, and should also <strong>in</strong>clude<br />

scientifi c and nonscientifi c considerations (Gomez-Pompa<br />

and Kaus, 1992; Westra, 1995; Light and Higgs, 1996;<br />

Higgs, 1997; Chauhan, 2003). Regier (1995) proposes an<br />

abstract defi nition for restoration that is dependent on<br />

what people believe as foster<strong>in</strong>g a state of ‘well-be<strong>in</strong>g.’<br />

Obviously, lack of conceptual clarity about restoration<br />

<strong>in</strong>troduces an element of uncerta<strong>in</strong>ty <strong>in</strong>to restoration<br />

problem solv<strong>in</strong>g. In this chapter, while be<strong>in</strong>g m<strong>in</strong>dful of<br />

<strong>the</strong> unresolved problems of conceptual clarity regard<strong>in</strong>g<br />

‘restoration’ o<strong>the</strong>r sources and implications of uncerta<strong>in</strong>ty<br />

and <strong>the</strong>ir relevance to river restoration are focused upon.<br />

1.2 BROAD PHILOSOPHICAL VIEWS ABOUT<br />

SCIENTIFIC UNCERTAINTY<br />

Dur<strong>in</strong>g <strong>the</strong> 19th century <strong>the</strong>re was a high degree of confi -<br />

dence <strong>in</strong> <strong>the</strong> methods and tools of science and technology<br />

to <strong>in</strong>crease understand<strong>in</strong>g of <strong>the</strong> natural world and enable<br />

robust predictions of its future states. This confi dence <strong>in</strong><br />

science contributed to beliefs that ‘nature’ could be controlled<br />

and rendered useful to humank<strong>in</strong>d (Latour, 1988).


Contribut<strong>in</strong>g to <strong>the</strong>se beliefs were philosophers and scientists<br />

(so-called ‘logical positivists’) who proposed that<br />

an important goal of science should focus on formulat<strong>in</strong>g<br />

hypo<strong>the</strong>ses and conduct<strong>in</strong>g observations to test <strong>the</strong>m,<br />

develop<strong>in</strong>g an understand<strong>in</strong>g of processes and l<strong>in</strong>kages<br />

among variables, and develop<strong>in</strong>g conclusions and predictions<br />

about which <strong>the</strong>re is a high degree of confi dence.<br />

More specifi cally, <strong>the</strong> logical positivistic view of science<br />

assumes that: knowledge is founded on experience; concepts<br />

and generalizations only represent <strong>the</strong> particulars<br />

from which <strong>the</strong>y have been abstracted; mean<strong>in</strong>g is<br />

grounded <strong>in</strong> observation; <strong>the</strong> sciences are unifi ed accord<strong>in</strong>g<br />

to <strong>the</strong> methodology of <strong>the</strong> natural sciences and <strong>the</strong><br />

ideal pursued <strong>in</strong> knowledge is <strong>the</strong> form of ma<strong>the</strong>matically<br />

formulated universal science deducible from <strong>the</strong> smallest<br />

number of possible axioms; and values are not facts<br />

grounded <strong>in</strong> observation and <strong>the</strong>refore cannot be <strong>in</strong>cluded<br />

as a part of scientifi c knowledge. One <strong>the</strong> one hand, while<br />

logical positivism has <strong>in</strong>fl uenced <strong>the</strong> th<strong>in</strong>k<strong>in</strong>g of modern<br />

scientists public policy makers, and decision makers, on<br />

<strong>the</strong> o<strong>the</strong>r it does not enjoy wide support from contemporary<br />

scientifi c philosophers (Hull, 1974).<br />

Scientists typically are conservative <strong>in</strong>sofar as <strong>the</strong>y provisionally<br />

reject a null hypo<strong>the</strong>sis only if <strong>the</strong> probability<br />

of mak<strong>in</strong>g a type I error is fi ve percent or less (Cranor,<br />

1993; Lemons et al., 1997). This scientifi c conservatism<br />

is consistent with <strong>the</strong> logical positivist goal of develop<strong>in</strong>g<br />

conclusions about which <strong>the</strong>re is a high degree of confi -<br />

dence. With respect to <strong>the</strong> use of science as a basis for<br />

public policy and decision mak<strong>in</strong>g, <strong>the</strong>re are those who<br />

hold that scientifi c methods and tools are capable of yield<strong>in</strong>g<br />

<strong>in</strong>formation about which <strong>the</strong>re is a high degree of scientifi<br />

c confi dence and, <strong>the</strong>refore, it is this <strong>in</strong>formation and<br />

not more speculative <strong>in</strong>formation that should be used as<br />

<strong>the</strong> basis for policy and decision mak<strong>in</strong>g (Peters, 1991;<br />

Sunste<strong>in</strong>, 2002). This latter view is a component of <strong>the</strong><br />

fi eld of environmental and human health risk assessment,<br />

which has developed to help <strong>in</strong>form public policy and<br />

decision makers about <strong>the</strong> risks from threats from both<br />

natural phenomena and human activities, <strong>in</strong>clud<strong>in</strong>g assess<strong>in</strong>g<br />

whe<strong>the</strong>r to undertake some river restoration projects.<br />

Components of risk analysis <strong>in</strong>clude: identify<strong>in</strong>g <strong>the</strong><br />

sequence of events through which exposure to risk could<br />

occur; determ<strong>in</strong><strong>in</strong>g <strong>the</strong> number and k<strong>in</strong>ds of people or<br />

environmental resources exposed to <strong>the</strong> risk; determ<strong>in</strong><strong>in</strong>g<br />

<strong>the</strong> adverse effects of exposure to <strong>the</strong> risks; and communicat<strong>in</strong>g<br />

risk assessment fi nd<strong>in</strong>gs to decision makers and<br />

<strong>the</strong> public. Although risk assessors acknowledge scientifi c<br />

uncerta<strong>in</strong>ty, <strong>the</strong>y often hold that scientifi c methods and<br />

tools can identify <strong>the</strong> risks and enable <strong>the</strong> calculation of<br />

<strong>the</strong> probabilities of <strong>the</strong>ir occurrence, <strong>in</strong>clud<strong>in</strong>g <strong>the</strong> bound<strong>in</strong>g<br />

of <strong>the</strong> probabilities with confi dence limits. For <strong>in</strong>-<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 5<br />

depth discussions on <strong>the</strong> role of scientifi c <strong>in</strong>formation <strong>in</strong><br />

policy and decision mak<strong>in</strong>g, see Peters (1991), Shrader-<br />

Frechette, (1994), Caldwell (1996), Lemons (1996), and<br />

Kaiser and Storvik (2003).<br />

Historically, logical positivism and its outgrowths also<br />

have <strong>in</strong>fl uenced <strong>the</strong> th<strong>in</strong>k<strong>in</strong>g of some scientists and policy<br />

makers <strong>in</strong> o<strong>the</strong>r ways by <strong>in</strong>culcat<strong>in</strong>g <strong>the</strong> view that ‘good’<br />

science is objective <strong>in</strong>sofar as it is not biased by <strong>the</strong> values<br />

of <strong>the</strong> scientists. Accord<strong>in</strong>gly, this view holds that <strong>the</strong><br />

proper role of science <strong>in</strong> policy and decision mak<strong>in</strong>g is to<br />

provide factual <strong>in</strong>formation to decision makers, and that<br />

any controversies about <strong>the</strong> factual <strong>in</strong>formation should be<br />

left to members of <strong>the</strong> scientifi c community competent<br />

<strong>in</strong> evaluat<strong>in</strong>g <strong>the</strong> scientifi c bases of <strong>the</strong> controversies<br />

(Shrader-Frechette, 1982). Consequently, <strong>the</strong> conclusions<br />

of scientifi c analyses do not become a part of broader<br />

public policy debates such as those that might perta<strong>in</strong> to<br />

such issues as what level of risk is acceptable. Practically<br />

speak<strong>in</strong>g, proponents of this view believe that <strong>the</strong> scientifi<br />

c and technical problems of manag<strong>in</strong>g large scale and<br />

complex problems are enormous and that <strong>the</strong> public cannot<br />

be expected to grasp <strong>the</strong> many scientifi c and technical<br />

issues <strong>in</strong>herent <strong>in</strong> understand<strong>in</strong>g and resolv<strong>in</strong>g <strong>the</strong> problems.<br />

Fur<strong>the</strong>r, <strong>the</strong> fundamental differences people have<br />

about how problems should be handled generate endless<br />

debate and controversy. This implies that while people and<br />

local governmental representatives with different <strong>in</strong>terests<br />

may review and comment on scientifi c and technical<br />

documents, <strong>the</strong>y would not be brought <strong>in</strong>to <strong>the</strong> actual<br />

decision- mak<strong>in</strong>g process regard<strong>in</strong>g <strong>the</strong> complex scientifi c<br />

dimensions of problems (Lemons et al., 1997).<br />

Despite <strong>the</strong> high degree of confi dence held by some<br />

people <strong>in</strong> scientifi c methods, confi dence <strong>in</strong> <strong>the</strong> power of<br />

science to understand and predict natural phenomena has<br />

been underm<strong>in</strong>ed by general relativity <strong>the</strong>ories, quantum<br />

<strong>the</strong>ories and chaos <strong>the</strong>ories (Brown, 1987). Rorty (1979)<br />

notes that <strong>the</strong>re is no evidence that science develops better<br />

and more accurate ‘mirrors’ with which to view nature. In<br />

his classic work, Kuhn (1962) describes how on <strong>the</strong> one<br />

hand <strong>the</strong> level of confi dence <strong>in</strong> models used by members<br />

of <strong>the</strong> scientifi c community <strong>in</strong>creases with evidence that<br />

supports <strong>the</strong> underly<strong>in</strong>g hypo<strong>the</strong>ses of <strong>the</strong> models, and on<br />

<strong>the</strong> o<strong>the</strong>r <strong>the</strong> scientists’ use of <strong>the</strong> models cannot be<br />

expected to produce consistently better and cumulatively<br />

more truthful descriptions of <strong>the</strong> way <strong>the</strong> world works.<br />

Accord<strong>in</strong>g to Kuhn, <strong>the</strong> reason is because predictive successes<br />

of scientifi c <strong>the</strong>ories do not guarantee <strong>the</strong>ir metaphysical<br />

accuracy because ‘paradigm shifts’ subsequently<br />

change scientists’ views of nature. O<strong>the</strong>r critics have<br />

po<strong>in</strong>ted out that so-called scientifi c truths of historical<br />

periods are social constructs <strong>in</strong>fl uenced by <strong>the</strong> dom<strong>in</strong>ant<br />

cultural and political powers of those periods (Briggs and


6 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Peat, 1982; Funtowizc and Ravetz, 1995). Some postmodern<br />

critics argue that Western science has been permeated<br />

by a variety of biases (e.g., ‘free market’ economics and<br />

<strong>in</strong>dustrialism, racism, religion, patriarchy) that while<br />

serv<strong>in</strong>g powerful <strong>in</strong>terests have not led to <strong>the</strong> generation<br />

and use of more ‘objective’ or value-free scientifi c<br />

knowledge (Sirageld<strong>in</strong>, 2002).<br />

More practically speak<strong>in</strong>g, scientifi c <strong>in</strong>stitutions as well<br />

as <strong>in</strong>dividual scientists <strong>in</strong>creas<strong>in</strong>gly hold <strong>the</strong> view that<br />

scientifi c uncerta<strong>in</strong>ty regard<strong>in</strong>g environment and human<br />

health problems is so pervasive and value laden that many<br />

conclusions about <strong>the</strong> problems cannot be made with<br />

a high degree of scientifi c confi dence (Cranor, 1993;<br />

Shrader-Frechette and McCoy, 1993; Lemons and Brown,<br />

1995; Lemons, 1996; EEA, 2001; Kriebel et al., 2001;<br />

Tickner, 2002, 2003). This view is based on empirical<br />

studies focus<strong>in</strong>g on: exposure to radiation from nuclear<br />

facilities and nuclear waste; manag<strong>in</strong>g large-scale ecosystems<br />

such as <strong>the</strong> Florida Everglades, agricultural lands,<br />

mar<strong>in</strong>e and freshwater oil spills; biodiversity protection<br />

and management of biological reserves; ocean dump<strong>in</strong>g<br />

of sewage sludge; sulfur dioxide and protection of human<br />

lungs to remote lake restoration; antifoul<strong>in</strong>g pa<strong>in</strong>ts on<br />

ships (e.g. tributylt<strong>in</strong>); estuar<strong>in</strong>e eutrophication; protection<br />

and management of mar<strong>in</strong>e fi sheries; extrapolat<strong>in</strong>g<br />

from toxicological responses <strong>in</strong> laboratory systems to both<br />

human health and to <strong>the</strong> responses of natural systems;<br />

management of fresh water resources; benzene <strong>in</strong> occupational<br />

sett<strong>in</strong>gs; <strong>the</strong> use and health impacts of asbestos;<br />

risks from polychlor<strong>in</strong>ated biphenyls (PCBs); halocarbons<br />

and <strong>the</strong> ozone layer; diethylstilbestrol (DES) and longterm<br />

consequences of prenatal exposure; human health<br />

effects of lead <strong>in</strong> <strong>the</strong> environment; methyl tertiary-butyl<br />

e<strong>the</strong>r (MBTE) <strong>in</strong> petrol as a substitute for lead; chemical<br />

contam<strong>in</strong>ation <strong>in</strong> <strong>the</strong> Great Lakes; hormones as growth<br />

promoters <strong>in</strong> animals used for food; and global climate<br />

change.<br />

1.3 WHAT IS ‘GOOD’ SCIENCE UNDER<br />

CONDITIONS OF UNCERTAINTY?<br />

Here, <strong>the</strong> question discussed is: What is ‘good’ science<br />

when science is used <strong>in</strong> try<strong>in</strong>g to solve river restoration<br />

problems under conditions of scientifi c uncerta<strong>in</strong>ty?<br />

A traditional and commonly accepted goal of science is<br />

that <strong>the</strong> probabilities of add<strong>in</strong>g speculative <strong>in</strong>formation to<br />

<strong>the</strong> body of scientifi c knowledge should be m<strong>in</strong>imal (Hull,<br />

1974; Peters, 1991). For this reason, scientists typically<br />

are conservative <strong>in</strong>sofar as <strong>the</strong>y provisionally reject a null<br />

hypo<strong>the</strong>sis if <strong>the</strong>re is a fi ve percent or less chance of<br />

reject<strong>in</strong>g it when it is true; this criterion is known as a<br />

normal standard of scientifi c proof or so-called ‘n<strong>in</strong>ety-<br />

fi ve percent confi dence rule.’ With respect to <strong>the</strong> science<br />

used to <strong>in</strong>form certa<strong>in</strong> types of river restoration policies<br />

and decisions, an example of a null hypo<strong>the</strong>sis is that <strong>the</strong>re<br />

is no effect on rivers or <strong>the</strong>ir resources from exist<strong>in</strong>g or<br />

proposed human activities. A type I error is to accept a<br />

false positive result, that is, to conclude that <strong>the</strong>re is harm<br />

to rivers or <strong>the</strong>ir resources when <strong>in</strong> fact <strong>the</strong>re is none. A<br />

type II error is to accept a false negative result, that is, to<br />

conclude <strong>the</strong>re is no harm when <strong>in</strong> fact <strong>the</strong>re is.<br />

Many environmental laws and regulations place <strong>the</strong><br />

burden of proof for demonstrat<strong>in</strong>g harm to <strong>the</strong> environment<br />

or human health on government regulatory agencies<br />

or o<strong>the</strong>rs attempt<strong>in</strong>g to demonstrate harm from development<br />

activities and, often, <strong>the</strong> standard that is used to meet<br />

<strong>the</strong> burden of proof test is <strong>the</strong> normal standard of scientifi c<br />

proof (Brown, 1995). When this standard is adopted as a<br />

basis for environmental decisions <strong>the</strong> scientifi c uncerta<strong>in</strong>ty<br />

that pervades many environmental problems means<br />

that <strong>the</strong> burden of proof usually will not be met, despite<br />

<strong>the</strong> fact that some <strong>in</strong>formation or even <strong>the</strong> weight of evidence<br />

might <strong>in</strong>dicate <strong>the</strong> existence of harm to <strong>the</strong> environment<br />

or human health. Consequently, <strong>in</strong> public policy and<br />

decision mak<strong>in</strong>g if <strong>the</strong> data show that some factor or perturbation<br />

has had an effect on <strong>the</strong> environment or human<br />

health but, say, only at <strong>the</strong> 70–90% confi dence level <strong>the</strong><br />

null hypo<strong>the</strong>sis that <strong>the</strong>re is no effect from <strong>the</strong> factor or<br />

perturbation is accepted. In such cases <strong>the</strong>re is a tendency<br />

by decision makers and o<strong>the</strong>rs to assume not only that<br />

<strong>the</strong>re was not enough evidence to reject <strong>the</strong> null hypo<strong>the</strong>sis<br />

but that <strong>the</strong>re was no effect when, <strong>in</strong> fact, <strong>the</strong> experimental<br />

design or test could have been too weak or <strong>the</strong> data too<br />

variable or too close for an effect to be demonstrated even<br />

if <strong>the</strong>re had been one (a type II error).<br />

M<strong>in</strong>imiz<strong>in</strong>g a type II error requires <strong>the</strong> statistical power<br />

of a research design or hypo<strong>the</strong>sis test to be calculated. In<br />

contrast to confi dence, which is designed to m<strong>in</strong>imize type<br />

I error, power depends on <strong>the</strong> magnitude of <strong>the</strong> hypo<strong>the</strong>sized<br />

change to be detected, <strong>the</strong> sample variance, <strong>the</strong><br />

number of replicates and <strong>the</strong> signifi cance value. The power<br />

of a test is <strong>the</strong> probability of reject<strong>in</strong>g a null hypo<strong>the</strong>sis<br />

when it is <strong>in</strong> fact false and should be rejected. The larger<br />

<strong>the</strong> detected change, <strong>the</strong> larger is <strong>the</strong> power. In situations<br />

where <strong>the</strong> detected changes are relatively small, statistical<br />

power is <strong>in</strong>creased by <strong>in</strong>creased sampl<strong>in</strong>g size but this<br />

<strong>in</strong>volves additional costs, research facilities and time.<br />

Analysis of variance <strong>in</strong> assess<strong>in</strong>g threats to environmental<br />

and human health problems shows that <strong>the</strong> number of<br />

samples required to yield a power of 0.95 <strong>in</strong>creases rapidly<br />

if changes smaller than 50% of <strong>the</strong> standard deviation are<br />

to be detected (Cranor, 1993). If <strong>the</strong> sample size stays<br />

<strong>the</strong> same <strong>the</strong> probability of a type I error is <strong>in</strong>creased if<br />

<strong>the</strong> probability of a type II error is decreased. A practical


problem <strong>in</strong> river restoration is that a desired emphasis on<br />

avoid<strong>in</strong>g type II error must be balanced aga<strong>in</strong>st o<strong>the</strong>r<br />

opportunities to use limited scientifi c resources to address<br />

o<strong>the</strong>r environmental and human health problems.<br />

Decisions about water management <strong>in</strong> <strong>the</strong> Klamath<br />

Bas<strong>in</strong> along <strong>the</strong> California and Oregon border <strong>in</strong> <strong>the</strong><br />

United States show some of <strong>the</strong> types of consequences that<br />

can happen when <strong>the</strong> law or decision makers require <strong>the</strong><br />

use of scientifi c <strong>in</strong>formation that meets <strong>the</strong> normal standard<br />

of scientifi c proof. In decades long disputes about<br />

water management <strong>in</strong> <strong>the</strong> bas<strong>in</strong>, federal biologists have<br />

been try<strong>in</strong>g to save three species of endangered fi sh by<br />

call<strong>in</strong>g for diversions of water from irrigation <strong>in</strong>to <strong>the</strong><br />

bas<strong>in</strong> to reduce <strong>the</strong> frequency of fi sh kills dur<strong>in</strong>g low water<br />

periods (over 30 000 Ch<strong>in</strong>ook salmon died dur<strong>in</strong>g a fi sh<br />

kill <strong>in</strong> 2002) (Service, 2003). As would be expected, a<br />

recommendation to reduce <strong>the</strong> amount of water available<br />

for irrigation met with strong opposition by ranchers and<br />

farmers <strong>in</strong> <strong>the</strong> bas<strong>in</strong>. However, failure of <strong>the</strong> biologists to<br />

meet normal scientifi c standards of proof demonstrat<strong>in</strong>g<br />

that releas<strong>in</strong>g more water <strong>in</strong>to <strong>the</strong> bas<strong>in</strong> would help <strong>the</strong><br />

fi sh has been cited by <strong>the</strong> United States Department of<br />

Interior (DOI) <strong>in</strong> its recent refusal to restrict <strong>the</strong> amount<br />

of water farmers can remove from waterways <strong>in</strong> <strong>the</strong> bas<strong>in</strong><br />

(NRC, 2004). It is important to understand that <strong>the</strong> DOI<br />

was not criticiz<strong>in</strong>g <strong>the</strong> scientists for do<strong>in</strong>g poor science;<br />

ra<strong>the</strong>r, it concluded that <strong>the</strong> normal standard of proof was<br />

not met. The DOI noted that factors such as nutrient runoff<br />

from natural sources as well as farms and ranches, algae<br />

blooms and dams that restrict access to fi shes’ spawn<strong>in</strong>g<br />

grounds complicate and <strong>in</strong> fact might preclude demonstrat<strong>in</strong>g<br />

<strong>the</strong> relation of water fl ow <strong>in</strong>to <strong>the</strong> bas<strong>in</strong> and <strong>the</strong><br />

health of <strong>the</strong> fi sh populations with a higher degree of<br />

scientifi c confi dence.<br />

The question of how to protect endangered species <strong>in</strong><br />

<strong>the</strong> Klamath Bas<strong>in</strong> and manage water resources raises a<br />

fundamental dilemma that those <strong>in</strong>volved <strong>in</strong> river restoration<br />

have to confront. On <strong>the</strong> one hand, traditional scientifi<br />

c norms call for mak<strong>in</strong>g conclusions on <strong>in</strong>formation<br />

about which <strong>the</strong>re is a high degree of confi dence. In <strong>the</strong><br />

Klamath Bas<strong>in</strong> example, adher<strong>in</strong>g to traditional scientifi c<br />

norms constra<strong>in</strong>s decisions to protect endangered fi sh<br />

under conditions of uncerta<strong>in</strong>ty but, at <strong>the</strong> same time, <strong>in</strong><br />

<strong>the</strong> absence of decisions to protect endangered fi sh <strong>the</strong><br />

threats cont<strong>in</strong>ue. In this type of situation, when science is<br />

used for public policy and decision mak<strong>in</strong>g, scientists<br />

might wish to consider whe<strong>the</strong>r and to what extent <strong>the</strong>y<br />

should be more comfortable with mak<strong>in</strong>g conclusions<br />

based on <strong>the</strong> weight of evidence ra<strong>the</strong>r than based solely<br />

or primarily on high levels of confi dence, especially s<strong>in</strong>ce<br />

public policy decisions are not based simply upon probabilistic<br />

considerations but ra<strong>the</strong>r <strong>in</strong>volve mak<strong>in</strong>g discrete<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 7<br />

and explicit choices among specifi c alternatives, <strong>in</strong>clud<strong>in</strong>g<br />

those with political, economic and ethical ramifi cations<br />

(Bella et al., 1994; Lemons et al., 1997). Admittedly, this<br />

could create a tension between do<strong>in</strong>g ‘good’ science as<br />

traditionally defi ned because scientists would be mak<strong>in</strong>g<br />

more speculative conclusions; however, <strong>in</strong> <strong>the</strong>ir attempt to<br />

make science rigorous <strong>in</strong> <strong>the</strong> sense of not want<strong>in</strong>g to add<br />

speculation to <strong>the</strong> body of scientifi c knowledge as required<br />

by <strong>the</strong> scientifi c profession <strong>the</strong> regulatory questions for<br />

which <strong>the</strong> studies are done may be frustrated.<br />

1.4 VALUE-LADEN DIMENSIONS OF SCIENCE<br />

AND UNCERTAINTY<br />

In addition to <strong>the</strong> policy and management problems that<br />

arise from <strong>the</strong> use of traditional scientifi c norms for<br />

mak<strong>in</strong>g conclusions <strong>in</strong> river restoration, o<strong>the</strong>r value-laden<br />

dimensions of science and policy both contribute to uncerta<strong>in</strong>ty<br />

and raise complicated questions about how it should<br />

be handled <strong>in</strong> public policy.<br />

Westra and Lemons (1995) and Lemons (1996) conta<strong>in</strong><br />

papers analyz<strong>in</strong>g both philosophical and scientifi c concepts<br />

used to <strong>in</strong>form ecological restoration science and<br />

practice. The concepts are diverse and <strong>in</strong>clude bas<strong>in</strong>g restoration<br />

on: ecosystems’ abilities to function successfully<br />

<strong>in</strong> a way deemed satisfactory by society; ecosystems’<br />

abilities to ma<strong>in</strong>ta<strong>in</strong> a balanced, <strong>in</strong>tegrated, adaptive<br />

community of organisms hav<strong>in</strong>g species composition,<br />

diversity and functional organization comparable to that<br />

of ‘natural’ habits of <strong>the</strong> region; ecosystems’ abilities to<br />

regenerate <strong>the</strong>mselves and withstand anthropogenic stress;<br />

and ecosystems’ abilities to approach optimum capacity<br />

for ecological succession development options. One<br />

problem with all <strong>the</strong>se defi nitions is that <strong>the</strong>y are <strong>in</strong>complete,<br />

general and qualitative <strong>in</strong>sofar as <strong>the</strong>y fail to provide<br />

precise pr<strong>in</strong>ciples that would make <strong>the</strong>m operational.<br />

In his analysis of value-laden issues <strong>in</strong> restoration<br />

for ecological as opposed to primarily or exclusively<br />

economic development goals, Cairns (2003) focuses on<br />

several types of problems. Firstly, some restoration projects<br />

are carried out on habitats different <strong>in</strong> k<strong>in</strong>d from those<br />

altered or destroyed. For example, an upland forest may<br />

be destroyed <strong>in</strong> order to partially restore river systems and<br />

wetlands that once occupied a particular lowland area.<br />

Despite <strong>the</strong> fact that restoration of rivers and/or wetlands<br />

has ecological value, sacrifi c<strong>in</strong>g a relatively undamaged<br />

habitat to restore ano<strong>the</strong>r k<strong>in</strong>d may cause unanticipated<br />

ecological change or harm. Secondly, with few exceptions<br />

most river and o<strong>the</strong>r ecological restoration projects are<br />

done to support <strong>the</strong> anthropocentric commodity or utilitarian<br />

values <strong>the</strong>y offer humans and this poses confl icts with<br />

restoration goals for nonanthropcentric reasons. Thirdly,


8 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

river restoration has uncerta<strong>in</strong> outcomes because of unpredictable<br />

events like fl oods or droughts, and because of <strong>the</strong><br />

limitations of <strong>the</strong> methods and tools of science to predict<br />

long-term outcomes. Fourthly, restoration efforts focus<strong>in</strong>g<br />

on s<strong>in</strong>gle species or ecosystem attributes might elim<strong>in</strong>ate<br />

those species that had <strong>in</strong>itially colonized disturbed areas<br />

and were at <strong>the</strong> same time able to tolerate anthropocentric<br />

stress. However, restoration projects might result <strong>in</strong> <strong>the</strong><br />

displacement of species tolerant to human activities with<br />

those less tolerant, at least <strong>in</strong> <strong>the</strong> short term. Fifthly, ecological<br />

restoration often takes place with species that<br />

tolerate anthropocentric stress and <strong>the</strong> ultimate succession<br />

processes and states will be human dom<strong>in</strong>ated or dependent.<br />

Most likely, a return to <strong>in</strong>digenous species would<br />

require cont<strong>in</strong>ual <strong>in</strong>tervention by researchers and<br />

environmental decision makers on behalf of <strong>the</strong>ir reestablishment.<br />

While science is not determ<strong>in</strong>ative to how<br />

<strong>the</strong> issues are resolved, robust scientifi c <strong>in</strong>formation is<br />

needed to help <strong>in</strong>form satisfactory policy judgments.<br />

Mayo and Hollander (1991), Cranor (1993), Shrader-<br />

Frechette and McCoy (1993) and Lemons and Brown<br />

(1995) analyzed how and why numerous value-laden<br />

judgments, evaluations, assumptions and <strong>in</strong>ferences are<br />

embedded <strong>in</strong> scientifi c methods perta<strong>in</strong><strong>in</strong>g to <strong>the</strong> study<br />

and management of ecosystems, <strong>in</strong>clud<strong>in</strong>g geohydrological<br />

and o<strong>the</strong>r water resources. For example, people have<br />

to decide <strong>the</strong> ecosystem parameters that are more important<br />

to base judgments on, often with little or no empirical<br />

<strong>in</strong>formation available. Assumptions have to be made, often<br />

without direct empirical evidence, whe<strong>the</strong>r ecosystem<br />

parameters should be considered <strong>in</strong>dependently or synergistically,<br />

and whe<strong>the</strong>r threshold values for environmental<br />

or health impacts exist and, if so, what such values should<br />

be. In addition, a lack of empirical data cannot be separated<br />

entirely from practical limitations imposed on environmental<br />

scientists. Decision makers require <strong>in</strong>formation<br />

<strong>in</strong> a relatively short period and at reasonable cost. These<br />

factors constra<strong>in</strong> <strong>the</strong> focus of most restoration studies to<br />

<strong>the</strong> short term, relatively small spatial areas and measurement<br />

of a relatively small number of samples and parameters.<br />

Fur<strong>the</strong>r, <strong>the</strong> above commentators conclude that<br />

many of <strong>the</strong> value-laden dimensions of scientifi c methodology<br />

and <strong>in</strong>formation not only are not fully recognized<br />

by scientists, policy and decision makers, but that <strong>the</strong><br />

failure to suffi ciently recognize <strong>the</strong> value-laden dimensions<br />

of science casts serious doubts about even <strong>the</strong> best<br />

and most thorough scientifi c and technical studies used to<br />

<strong>in</strong>form decisions about problems such as river restoration.<br />

In o<strong>the</strong>r words, unless <strong>the</strong> value-laden dimensions of scientifi<br />

c studies are disclosed <strong>the</strong> positions of decision<br />

makers will appear to be justifi ed on value-neutral scientifi<br />

c reason<strong>in</strong>g and will appear to be more certa<strong>in</strong> than<br />

warranted when, <strong>in</strong> fact, <strong>the</strong> positions will be based, <strong>in</strong><br />

part, on often controversial and confl ict<strong>in</strong>g values of scientists<br />

and decision makers (see also Fleck, 1979).<br />

One of <strong>the</strong> most common ways <strong>in</strong> which value issues<br />

are hidden <strong>in</strong> public policy concern<strong>in</strong>g issues such as river<br />

restoration develops out of <strong>the</strong> expectation that technical<br />

analysts can isolate and apply <strong>the</strong> facts under dispute <strong>in</strong><br />

a manner consistent with policy directives or legislative<br />

mandates. This separation of facts and values is highly<br />

problematic. For example, consider <strong>the</strong> use of safety<br />

factors <strong>in</strong> river water quality regulations as a means of<br />

extra protection for human or environmental health.<br />

Implicit <strong>in</strong> <strong>the</strong> choice of safety factors is an asymmetric<br />

cost function with health costs ris<strong>in</strong>g more steeply than<br />

costs for over-treatment. Implicit <strong>in</strong> <strong>the</strong> magnitude of a<br />

safety factor are signifi cant uncerta<strong>in</strong>ties <strong>in</strong> health impacts<br />

and a steeper cost function for health effects from undertreatment<br />

than for over-treatment. When <strong>the</strong>se issues<br />

rema<strong>in</strong> implicit <strong>in</strong> <strong>the</strong> use of safety factors (as <strong>the</strong>y typically<br />

are) <strong>the</strong> real issues of knowledge and uncerta<strong>in</strong>ty are<br />

obscured for decision makers and <strong>the</strong> public. Often, <strong>the</strong>se<br />

issues rema<strong>in</strong> implicit or hidden because safety factors<br />

and cost factors are described <strong>in</strong> quantitative terms perta<strong>in</strong><strong>in</strong>g<br />

to risks or cost–benefi t calculations. This <strong>in</strong>creases<br />

<strong>the</strong> likelihood of <strong>the</strong> misuse of conclusions by decision<br />

makers who do not understand <strong>the</strong> basis for deriv<strong>in</strong>g safety<br />

factors (Brown, 1987).<br />

1.5 PRACTICAL AND POLICY ASPECTS<br />

OF UNCERTAINTY<br />

Cairns (2001) analyzed how most complex environmental<br />

problems transcend <strong>the</strong> capabilities of any s<strong>in</strong>gle discipl<strong>in</strong>e<br />

but at <strong>the</strong> same time and all too often research teams<br />

are not suffi ciently <strong>in</strong>terdiscipl<strong>in</strong>ary to deal adequately<br />

with <strong>the</strong> problems. In addition, problem solv<strong>in</strong>g often does<br />

not provide a balanced mix of academicians, public policy<br />

and decision makers, representatives from private <strong>in</strong>dustry<br />

or bus<strong>in</strong>ess and nongovernmental organizations. As a<br />

result, <strong>the</strong> fram<strong>in</strong>g of problems and <strong>the</strong>ir solution is too<br />

often fragmented and <strong>in</strong>effectual and biased towards one<br />

or a few discipl<strong>in</strong>ary approaches or stakeholder groups<br />

(Nienhuis and Leuven, 2001; Benyam<strong>in</strong>e, 2002).<br />

Some scientists and policy makers <strong>in</strong>volved <strong>in</strong> environmental<br />

problem solv<strong>in</strong>g have argued for syn<strong>the</strong>siz<strong>in</strong>g<br />

analyses and alternatives to solutions of environmental<br />

resource problems (Lubchenco et al., 1991; Bella et al.,<br />

1994; Lemons and Brown, 1995; Caldwell, 1996). In practice,<br />

at least three levels of syn<strong>the</strong>sis may be identifi ed.<br />

The fi rst is conceptual syn<strong>the</strong>sis and occurs when <strong>the</strong><br />

diverse and often disparate elements of a problem situation<br />

are pulled toge<strong>the</strong>r <strong>in</strong>tuitively, <strong>the</strong>n tested and <strong>in</strong>tegrated


to form a coherent research design. Follow<strong>in</strong>g analysis of<br />

<strong>the</strong> problem and identifi cation of its causes and consequences,<br />

a second level of syn<strong>the</strong>sis <strong>in</strong>volves del<strong>in</strong>eation<br />

of <strong>the</strong> fi nd<strong>in</strong>gs of <strong>the</strong> scientifi c research. A third level of<br />

syn<strong>the</strong>sis can occur when research fi nd<strong>in</strong>gs are evaluated<br />

and consolidated <strong>in</strong> decid<strong>in</strong>g a course of action by<br />

decision makers.<br />

Despite <strong>the</strong> need for greater syn<strong>the</strong>sis of research<br />

methods and <strong>in</strong>formation, syn<strong>the</strong>sis itself <strong>in</strong>troduces additional<br />

value-laden dimensions and uncerta<strong>in</strong>ties <strong>in</strong>to environmental<br />

problem solv<strong>in</strong>g. Caldwell (1996) and Brown<br />

(1995) discuss how decision makers must syn<strong>the</strong>size a<br />

policy (<strong>in</strong> part) from <strong>the</strong> scientifi c <strong>in</strong>formation available<br />

even when <strong>the</strong> <strong>in</strong>formation often is <strong>in</strong>complete. When<br />

science is used to <strong>in</strong>form policy decisions such decisions<br />

also <strong>in</strong>clude economic, legal, adm<strong>in</strong>istrative and cultural<br />

parameters and, <strong>the</strong>refore, are based on human values<br />

and judgments. Benyam<strong>in</strong>e (2002) discusses how disagreements<br />

about scientifi c <strong>the</strong>ories that are used as a<br />

basis for <strong>in</strong>form<strong>in</strong>g public policy and decision mak<strong>in</strong>g<br />

become entangled with economic, legal and ideological<br />

issues. Sometimes, <strong>the</strong> disagreements rema<strong>in</strong> largely confi<br />

ned to <strong>the</strong> scientifi c community, while at o<strong>the</strong>r times <strong>the</strong><br />

public knows about <strong>the</strong>m. When scientists and/or decision<br />

makers know <strong>the</strong> underly<strong>in</strong>g <strong>the</strong>oretical bases for disagreements,<br />

this knowledge can <strong>in</strong>fl uence <strong>the</strong> scientifi c<br />

arguments about <strong>the</strong> disagreements. However, some confl<br />

ict<strong>in</strong>g arguments and <strong>the</strong>ir underly<strong>in</strong>g <strong>the</strong>oretical support<br />

can be under recognized or little understood by <strong>the</strong> nonscientifi<br />

c communities as well as by scientists whose specialized<br />

fi elds are outside <strong>the</strong> discipl<strong>in</strong>e where debates<br />

about <strong>the</strong>ories are tak<strong>in</strong>g place. When this happens, confl<br />

ict<strong>in</strong>g scientifi c arguments will not have much <strong>in</strong>fl uence<br />

on <strong>the</strong> disagreements.<br />

There is debate with<strong>in</strong> <strong>the</strong> scientifi c and public policy<br />

communities regard<strong>in</strong>g approaches to deal with uncerta<strong>in</strong>ties<br />

(Bradshaw and Borchers, 2000). For example, one<br />

approach might be to attempt to <strong>in</strong>crease scientifi c confi -<br />

dence by <strong>in</strong>creas<strong>in</strong>g scientifi c confi rmation of hypo<strong>the</strong>ses.<br />

In this way, scientists can decrease uncerta<strong>in</strong>ty suffi ciently<br />

to allow more precise estimates of risk for policy and<br />

decision makers. A second approach might be to <strong>in</strong>crease<br />

<strong>the</strong> knowledge of sources of uncerta<strong>in</strong>ty by enhanc<strong>in</strong>g<br />

education and communication between scientists, policy<br />

and decision makers and <strong>the</strong> general public. A benefi t of<br />

this approach is that when scientists and decision makers<br />

are <strong>in</strong>volved with <strong>the</strong> public <strong>the</strong>re is greater opportunity<br />

for consensus build<strong>in</strong>g and less risk of legal challenges<br />

from disaffected stakeholders. A third approach might be<br />

to foster <strong>the</strong> view that scientifi c uncerta<strong>in</strong>ty should be<br />

regarded <strong>in</strong> public policy and decision mak<strong>in</strong>g as it is<br />

with<strong>in</strong> <strong>the</strong> scientifi c community, namely, as <strong>in</strong>formation<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 9<br />

for hypo<strong>the</strong>sis build<strong>in</strong>g and test<strong>in</strong>g. Consequently, calls<br />

for faster and more ‘certa<strong>in</strong>’ scientifi c conclusions to<br />

<strong>in</strong>form public policy and decision mak<strong>in</strong>g would be tempered<br />

with a better understand<strong>in</strong>g of <strong>the</strong> limitations and<br />

capabilities of science to provide <strong>in</strong>formation about which<br />

<strong>the</strong>re is a high degree of confi dence.<br />

Still ano<strong>the</strong>r approach might be for society to require<br />

procedural rules for mak<strong>in</strong>g decisions under conditions<br />

of scientifi c uncerta<strong>in</strong>ty to take <strong>in</strong>to account confl ict<strong>in</strong>g<br />

po<strong>in</strong>ts of view, possible consequences to welfare, as well<br />

as various ethical and legal obligations such as those<br />

<strong>in</strong>volv<strong>in</strong>g free <strong>in</strong>formed consent and due process (Shrader-<br />

Frechette, 1996). This approach could <strong>in</strong>clude greater use<br />

of <strong>the</strong> precautionary pr<strong>in</strong>ciple by help<strong>in</strong>g to ensure that<br />

when <strong>the</strong>re is substantial scientifi c uncerta<strong>in</strong>ty about <strong>the</strong><br />

risks and benefi ts of a proposed activity, policy decisions<br />

should be made <strong>in</strong> a way that errs on <strong>the</strong> side of caution<br />

with respect to <strong>the</strong> environment and <strong>the</strong> health of <strong>the</strong><br />

public (Kriebel et al., 2001; Tickner, 2003).<br />

1.6 CASE STUDY OF SCIENTIFIC<br />

UNCERTAINTY IN RIVER RESTORATION<br />

The example discussed here is based on ecological studies<br />

conducted from 1980–1989 <strong>in</strong> a small (4th order), black<br />

water West African river, <strong>the</strong> <strong>River</strong> Ikpoba fl ow<strong>in</strong>g through<br />

Ben<strong>in</strong> City, Sou<strong>the</strong>rn Nigeria (Victor and Dickson, 1985;<br />

Victor and Ogbeibu, 1985, 1986, 1991; Victor and Tetteh,<br />

1988; Ogbeibu and Victor, 1989; Victor and Brown, 1990;<br />

Victor and Meye, 1994; Victor and Onomivbori, 1996;<br />

Victor, 1998). The stretch of river studied was affected by<br />

a variety of urban perturbations such as damm<strong>in</strong>g, water<br />

extraction, po<strong>in</strong>t and nonpo<strong>in</strong>t source pollution, sand<br />

dredg<strong>in</strong>g and agriculture. As a result of government<br />

policies and directives mandat<strong>in</strong>g river clean-up activities,<br />

<strong>the</strong>re was a rare opportunity to study river restoration by<br />

recovery processes. Scientifi c results of this study were<br />

published <strong>in</strong> <strong>the</strong> series of publications listed above and<br />

provide one of <strong>the</strong> bases of our focus on uncerta<strong>in</strong>ties<br />

associated with <strong>the</strong> restoration process.<br />

The fi rst logical step was to <strong>in</strong>vestigate recovery processes.<br />

Geomorphologic changes of <strong>the</strong> river channel and<br />

<strong>the</strong> entire riparian corridor <strong>in</strong>fl uenced by urban development<br />

could not be reversed (e.g. <strong>the</strong> presence of a dam,<br />

water extraction for human consumption) and <strong>the</strong>refore<br />

complete restoration would not be possible. Removal of<br />

human <strong>in</strong>fl uences where possible would permit recovery,<br />

but <strong>the</strong> rates limit<strong>in</strong>g recovery <strong>in</strong> different sections would<br />

not only depend on <strong>the</strong> type of <strong>in</strong>fl uence (e.g. sand extraction,<br />

car wash<strong>in</strong>g), but would also be complicated by<br />

natural events such as fl oods. Thus <strong>the</strong> optimum threshold<br />

for <strong>the</strong> recovery process <strong>in</strong> this study at various sections


10 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

of <strong>the</strong> river cont<strong>in</strong>uum was unpredictable and uncerta<strong>in</strong>.<br />

O<strong>the</strong>r signifi cant uncerta<strong>in</strong>ties were: <strong>the</strong> role of early<br />

recoloniz<strong>in</strong>g species affect<strong>in</strong>g <strong>the</strong> trajectory of recovery;<br />

<strong>the</strong> successional sequence of species re-establish<strong>in</strong>g; and<br />

<strong>the</strong> establishment of appropriate abiotic conditions and<br />

<strong>the</strong> establishment of previously non-exist<strong>in</strong>g non-native<br />

species like <strong>the</strong> water hyac<strong>in</strong>th.<br />

The next group of uncerta<strong>in</strong>ties was related to <strong>the</strong> analysis<br />

and syn<strong>the</strong>sis of data. Removal of a particular human<br />

<strong>in</strong>fl uence (e.g. discharge of untreated sewage) <strong>in</strong> one<br />

section signifi cantly <strong>in</strong>creased <strong>the</strong> presence of a parameter,<br />

say i (P < 0.05), show<strong>in</strong>g that this parameter was a good<br />

<strong>in</strong>dicator of recovery. But <strong>the</strong> same parameter did not<br />

<strong>in</strong>crease signifi cantly <strong>in</strong> an adjacent section with a similar<br />

problem (P > 0.05) show<strong>in</strong>g its uncerta<strong>in</strong> predictive status.<br />

Graphical exam<strong>in</strong>ation of associations between specifi c<br />

human <strong>in</strong>fl uences (e.g. removal of detergent contam<strong>in</strong>ation)<br />

and biological parameters like taxa richness and<br />

abundance showed positive relationships, but statistically<br />

<strong>the</strong>se relationships as evaluated by Pearson’s r or Spearman’s<br />

r s were not signifi cant (P > 0.05). Thus, correlation<br />

matrices generated for evaluat<strong>in</strong>g relationships between<br />

<strong>the</strong> removal of perturbation <strong>in</strong>fl uences and <strong>the</strong> recovery of<br />

both biotic and abiotic parameters were diffi cult to <strong>in</strong>terpret.<br />

Interpretation us<strong>in</strong>g traditional statistical norms and<br />

acceptable levels of signifi cance were ecologically and<br />

rationally highly problematic.<br />

Fur<strong>the</strong>r uncerta<strong>in</strong>ties arose while consider<strong>in</strong>g <strong>the</strong> temporal<br />

and spatial scale of <strong>the</strong> recovery process. The recovery<br />

process was happen<strong>in</strong>g <strong>in</strong> an urban sett<strong>in</strong>g with a new<br />

land use matrix, far different from prist<strong>in</strong>e or semiprist<strong>in</strong>e<br />

natural conditions that previously existed. Therefore, comparison<br />

of <strong>the</strong> restored river sections to that of ‘undisturbed’<br />

sections upstream was not valid and new basel<strong>in</strong>e<br />

standards had to be established for future monitor<strong>in</strong>g.<br />

Even <strong>the</strong>se were extremely site specifi c with very limited<br />

potential for use <strong>in</strong> o<strong>the</strong>r sections of <strong>the</strong> study stretch.<br />

Because of <strong>the</strong> uncerta<strong>in</strong>ties <strong>in</strong>volved, <strong>the</strong> scale needed for<br />

manag<strong>in</strong>g temporal and spatial variability <strong>in</strong> restoration<br />

was not apparent. ‘Rules of thumb’ based on value judgments<br />

had to be made to evaluate recovery <strong>in</strong> specifi c<br />

sections of <strong>the</strong> river stretch with specifi c types of perturbations.<br />

The magnitude of uncerta<strong>in</strong>ties <strong>in</strong>volved render <strong>the</strong><br />

comb<strong>in</strong>ation of tools used here (e.g. sampl<strong>in</strong>g duration,<br />

sampl<strong>in</strong>g frequencies, choice of methods, size of samples,<br />

analytical models) <strong>in</strong>adequate to evaluate recovery processes<br />

<strong>in</strong> o<strong>the</strong>r rivers of similar stream order, larger rivers<br />

with higher stream order and even <strong>the</strong> same river 100 km<br />

downstream where its stream order is >8.<br />

Implementation and analysis of monitor<strong>in</strong>g were also<br />

wrought with uncerta<strong>in</strong>ties. For example, fi ve different<br />

sections of <strong>the</strong> river stretch were monitored for restoration<br />

by recovery. Each section was characterized by its own set<br />

of physical and biological parameters that were good<br />

<strong>in</strong>dicators of recovery at <strong>the</strong> time of <strong>the</strong> study. Due to<br />

limitations of fund<strong>in</strong>g, personnel and <strong>the</strong> required cost<br />

effectiveness of <strong>the</strong> monitor<strong>in</strong>g program, proposals had<br />

to identify common parameters that would monitor <strong>the</strong><br />

overall health of <strong>the</strong> study stretch <strong>in</strong> <strong>the</strong> long term. As<br />

discussed earlier, uncerta<strong>in</strong>ties associated with <strong>the</strong> analysis<br />

and syn<strong>the</strong>sis of data did not permit <strong>the</strong> ready identifi<br />

cation of common parameters. Even if <strong>the</strong>re was an<br />

agreement on us<strong>in</strong>g different sets of parameters for different<br />

sections of <strong>the</strong> stretch, <strong>the</strong>re was no certa<strong>in</strong>ty that <strong>the</strong>se<br />

parameters (e.g. BOD, nitrate–N, fi sh diversity) will cont<strong>in</strong>ue<br />

to serve as good <strong>in</strong>dicators of recovery <strong>in</strong> <strong>the</strong> long<br />

term. It was also possible that a parameter considered<br />

trivial and not <strong>in</strong>cluded <strong>in</strong> <strong>the</strong> monitor<strong>in</strong>g program (e.g.<br />

dissolved organic matter, haptobenthos) may become<br />

important <strong>in</strong> <strong>the</strong> long term, which <strong>in</strong> itself cannot be<br />

defi ned clearly. ‘Long term’ <strong>in</strong> this case at least did not<br />

refer to an <strong>in</strong>defi nite period and envisaged monitor<strong>in</strong>g<br />

programs were not relatively open-ended, as often is <strong>the</strong><br />

case <strong>in</strong> countries with limited resources. Policy and decision<br />

makers considered what seemed to be a comprehensive<br />

proposal for monitor<strong>in</strong>g <strong>in</strong> <strong>the</strong> view of scientists as<br />

not be<strong>in</strong>g practical.<br />

Policy questions concern<strong>in</strong>g river restoration <strong>in</strong> <strong>the</strong><br />

geopolitical context were plagued with more uncerta<strong>in</strong>ties<br />

than scientifi c questions. The political climate of <strong>the</strong> study<br />

area at that time was unstable and government changed<br />

hands frequently. For example, one government downgraded<br />

<strong>the</strong> priority given to environmental issues, such as<br />

river restoration, by <strong>the</strong> previous government if personal<br />

<strong>in</strong>terests and political expediency demanded it. Assum<strong>in</strong>g<br />

no change <strong>in</strong> policies with change <strong>in</strong> governments, <strong>the</strong>re<br />

were uncerta<strong>in</strong>ties concern<strong>in</strong>g fund<strong>in</strong>g tools that would<br />

ensure <strong>the</strong> long term success of restoration, design of<br />

legislation to accommodate river restoration without compromis<strong>in</strong>g<br />

susta<strong>in</strong>able development and coord<strong>in</strong>ation of<br />

policies and legislation to devise strategies for river restoration<br />

<strong>in</strong> a broader context of <strong>the</strong> adm<strong>in</strong>istrative region<br />

(e.g. district, state, country). The management of restored<br />

or recovered river as a water resource for domestic use,<br />

agriculture, fi sheries and recreation was not considered<br />

<strong>in</strong>tentionally. For scientifi c uncerta<strong>in</strong>ty concern<strong>in</strong>g water<br />

resources management, see Canter (1996).<br />

1.7 CONCLUSION<br />

Scientifi c and o<strong>the</strong>r uncerta<strong>in</strong>ty is pervasive <strong>in</strong> environmental<br />

problem solv<strong>in</strong>g, and river restoration is no exception.<br />

When <strong>the</strong> traditional scientifi c standard of proof is<br />

used as a basis for river restoration decisions, <strong>the</strong> scientifi c


uncerta<strong>in</strong>ty that pervades many restoration problems<br />

means that <strong>the</strong> standard usually will not be met, despite<br />

<strong>the</strong> fact that some <strong>in</strong>formation or even <strong>the</strong> weight of evidence<br />

might <strong>in</strong>dicate <strong>the</strong> existence of harm and <strong>the</strong>refore<br />

<strong>the</strong> need for restoration. A high degree of confi dence <strong>in</strong><br />

river restoration science, as <strong>in</strong> o<strong>the</strong>r sciences, unfortunately<br />

seems to h<strong>in</strong>ge on conventional statistical decision<br />

rules such as when, for example, river monitor<strong>in</strong>g dur<strong>in</strong>g<br />

restoration strives to detect human-<strong>in</strong>fl uenced factors<br />

that caused deviations from basel<strong>in</strong>e conditions. The major<br />

concern here will be ecological change and not how large<br />

or small <strong>the</strong> P-values are (Yoccoz, 1991; Stewart-Oaten,<br />

1996). Most statistical decision rules are too simplistic and<br />

mislead<strong>in</strong>g <strong>in</strong>sofar as <strong>the</strong>ir assumptions that lack of statistical<br />

signifi cance means lack of environmental signifi -<br />

cance (Karr and Chu, 1999). Accord<strong>in</strong>g to Yoccoz (1991),<br />

Kriebel et al. (2001), and Lemons et al. (1997) ecologists<br />

tend to over-use tests of signifi cance and restoration ecologists<br />

are no exception to this rule. Karr and Chu (1999)<br />

suggest that it would be wiser to decide what is ecologically<br />

relevant fi rst and <strong>the</strong>n use hypo<strong>the</strong>sis test<strong>in</strong>g to detect<br />

ecologically relevant effects; <strong>the</strong> use of o<strong>the</strong>r statistical<br />

tools such as power analysis and decision <strong>the</strong>ory also is<br />

recommended (Hilborn, 1997).<br />

Cairns and Heckman (1996) state that restoration<br />

ecology <strong>in</strong> general ‘is a bridge between <strong>the</strong> social and<br />

natural sciences.’ In this chapter it has been shown that it<br />

is impossible to separate scientifi c and policy questions <strong>in</strong><br />

restoration ecology and this, <strong>in</strong> and of itself, <strong>in</strong>troduces<br />

uncerta<strong>in</strong>ty <strong>in</strong>to what o<strong>the</strong>rwise might be viewed as value–<br />

neutral or ‘objective’ scientifi c conclusions.<br />

As discussed more generally <strong>in</strong> this chapter and shown<br />

more specifi cally <strong>in</strong> <strong>the</strong> case study section, scientifi c<br />

research is both value-laden and is used to support politically-driven<br />

river restoration policies and decision mak<strong>in</strong>g<br />

(see also Shrader-Frechette, 1994). For example, historical<br />

or descriptive research is <strong>in</strong>tended to reveal or expla<strong>in</strong> <strong>the</strong><br />

dynamics of a given policy and to explore its orig<strong>in</strong> and<br />

evolution. Prescriptive or advocacy research defends a<br />

conclusion or possibly even a preconceived policy, and<br />

also is characterized by publicized disputes among, e.g.,<br />

scientists. Decision-<strong>in</strong>form<strong>in</strong>g or predictive research typically<br />

is fi nanced by grants or contracts lead<strong>in</strong>g to conclusions<br />

supportive of a predeterm<strong>in</strong>ed policy preference,<br />

sponsor bias, or predilections with<strong>in</strong> a research peer group.<br />

Consequently, <strong>the</strong> focus of this research does not attempt<br />

to analyze all feasible alternative policy choices and <strong>the</strong><br />

probable consequences. Because <strong>the</strong> focus of this research<br />

is on applicability for a particular policy its fi nd<strong>in</strong>gs are<br />

presented <strong>in</strong> <strong>the</strong> form of propositions upon which decisions<br />

can be made. The effi cacy of <strong>the</strong> policy towards<br />

which <strong>the</strong> research is focused depends on <strong>the</strong> validity,<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 11<br />

reliability and persuasiveness of <strong>the</strong> research and <strong>the</strong><br />

extent of political public receptivity.<br />

It is important to clearly dist<strong>in</strong>guish between <strong>the</strong> use of<br />

methods and tools of science to understand <strong>the</strong> phenomena<br />

of nature and <strong>the</strong> acquisition of scientifi c <strong>in</strong>formation<br />

about a restoration issue and <strong>the</strong> sett<strong>in</strong>g of policy; but <strong>in</strong><br />

practice, <strong>the</strong>re is not always an unambiguous demarcation.<br />

Policy makers set agendas that determ<strong>in</strong>e <strong>the</strong> questions<br />

that are asked of scientists; scientists formulate hypo<strong>the</strong>ses<br />

<strong>in</strong> ways limited by <strong>the</strong>ir tools and <strong>the</strong>ir imag<strong>in</strong>ations<br />

and discipl<strong>in</strong>ary conventions. Consequently, <strong>the</strong> <strong>in</strong>formation<br />

<strong>the</strong>y provide to <strong>the</strong> policy makers is limited and<br />

socially determ<strong>in</strong>ed to a degree and <strong>the</strong>refore <strong>the</strong>re is a<br />

complicated feedback relation between <strong>the</strong> discoveries of<br />

science and <strong>the</strong> sett<strong>in</strong>g of policy. While attempt<strong>in</strong>g to be<br />

objective and focus on understand<strong>in</strong>g river restoration<br />

phenomena, scientists and o<strong>the</strong>r researchers should be<br />

aware of <strong>the</strong> policy uses of <strong>the</strong>ir work and of <strong>the</strong>ir social<br />

responsibility to carryout science that protects <strong>the</strong> environment<br />

and human health (Kriebel et al., 2001). In try<strong>in</strong>g<br />

to fulfi ll this responsibility, scientifi c and o<strong>the</strong>r uncerta<strong>in</strong>ty<br />

needs to be taken <strong>in</strong>to greater account.<br />

The discussion of some of <strong>the</strong> value-laden decisions and<br />

judgments scientists and o<strong>the</strong>r researchers make is not a<br />

criticism. Ra<strong>the</strong>r, <strong>the</strong> issue is discussed because a failure to<br />

recognize <strong>the</strong> existence of <strong>the</strong> value-laden dimensions of<br />

science casts serious doubt about even <strong>the</strong> best and most<br />

thorough of scientifi c and technical studies used to <strong>in</strong>form<br />

decisions about river restoration. In o<strong>the</strong>r words, unless <strong>the</strong><br />

value-laden dimensions of scientifi c and technical studies<br />

used to derive <strong>in</strong>formation are disclosed, <strong>the</strong> positions<br />

of policy makers and decision makers will appear to be<br />

justifi ed on objective or value–neutral scientifi c reason<strong>in</strong>g<br />

when, <strong>in</strong> fact, <strong>the</strong>y will be based <strong>in</strong> part on often controversial<br />

or confl ict<strong>in</strong>g values of scientists <strong>the</strong>mselves.<br />

REFERENCES<br />

Balon EK, Coche AG. 1974. Lake Kariba, A Man-Made Tropical<br />

Ecosystem <strong>in</strong> Central Africa. DW Junk Publishers: The Hague,<br />

The Ne<strong>the</strong>rlands.<br />

Bella DA, Jacobs R, Hiram L. 1994. Ecological <strong>in</strong>dicators of<br />

global climate change: A research framework. Environmental<br />

Management 18: 489–500.<br />

Benyam<strong>in</strong>e M. 2002. Theoretical Disputes and Practical Environmental<br />

Dilemmas. Orebro Studies <strong>in</strong> Environmental Science<br />

3. Orebro University: Orebro.<br />

Bradshaw GA, Borchers JG. 2000. <strong>Uncerta<strong>in</strong>ty</strong> as <strong>in</strong>formation:<br />

Narrow<strong>in</strong>g <strong>the</strong> Science – Policy gap. Conservation Ecology<br />

4: 7. [onl<strong>in</strong>e] URL: http://www.consecol.org/vol4/iss1/art7/)<br />

(accessed 23 March 2004)<br />

Briggs J, Peat F. 1982. Look<strong>in</strong>g Glass Universe. Simon and<br />

Schuster: New York.


12 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Brown D. 1987. Ethics, science, and environmental regulation.<br />

Environmental Ethics 9: 331–350.<br />

Brown D. 1995. The role of ethics <strong>in</strong> susta<strong>in</strong>able development and<br />

environmental protection decisionmak<strong>in</strong>g. In Susta<strong>in</strong>able<br />

Development: Science, Ethics, and Public Policy, Lemons J,<br />

Brown D (Eds). Kluwer Academic Publisher: Dordrecht, The<br />

Ne<strong>the</strong>rlands; 39–51.<br />

Cairns (Jr) J. 2001. Rationale for restoration. In Handbook of<br />

Ecological <strong>Restoration</strong>, Vol. 1: Pr<strong>in</strong>ciples and <strong>Restoration</strong>,<br />

Davy AJ, Perry, J (Eds). Cambridge University Press:<br />

Cambridge; 10–23.<br />

Cairns (Jr) J. 2003. Ethical issues <strong>in</strong> ecological restoration. Ethics<br />

<strong>in</strong> Science and Environmental Politics 3: 50–61.<br />

Cairns J (Jr), Heckman JR. 1996. <strong>Restoration</strong> ecology: The<br />

state of an emerg<strong>in</strong>g fi eld. Annual Review of Energy and<br />

Environment 21: 167–187.<br />

Caldwell LK. 1996. In Scientifi c <strong>Uncerta<strong>in</strong>ty</strong> and Environmental<br />

Problem Solv<strong>in</strong>g, Lemons J (Ed). Blackwell Science,<br />

Cambridge; 394–422.<br />

Canter LW. 1996. Scientifi c uncerta<strong>in</strong>ty and water resources management.<br />

In Scientifi c <strong>Uncerta<strong>in</strong>ty</strong> and Environmental Problem<br />

Solv<strong>in</strong>g, Lemons J (Ed). Blackwell Science Inc.: Cambridge;<br />

264–297.<br />

Chauhan M. 2003. Conserv<strong>in</strong>g biodiversity <strong>in</strong> arid regions: Experiences<br />

with protected areas <strong>in</strong> India. In Conserv<strong>in</strong>g Biodiversity<br />

<strong>in</strong> Arid Regions. Lemons J, Victor R, Schaffer D (Eds).<br />

Kluwer Academic Publishers: Boston.<br />

Cranor C. 1993. Regulat<strong>in</strong>g Toxic Substances: A Philosophy of<br />

Science and Law. Oxford University Press, New York.<br />

De Wall LC, Large ARG, Gippel CJ, Wade PM. 1995. <strong>River</strong> and<br />

fl oodpla<strong>in</strong> rehabilitation <strong>in</strong> Western Europe: Opportunities and<br />

constra<strong>in</strong>ts. Archives Hydrobiologia Supplement 101 – Large<br />

<strong>River</strong>s 9: 679–693.<br />

Eiseltova M, Biggs J. 1995. <strong>Restoration</strong> of stream ecosystems: An<br />

<strong>in</strong>tegrated approach. International Waterfowl and Wetlands<br />

Research Bureau Publication 37: 1–170.<br />

(EEA) European Environment Agency. 2001. Late Lessons From<br />

Early Warn<strong>in</strong>gs: The Precautionary Pr<strong>in</strong>ciple 1896–2000,<br />

Environmental Issue Report No. 22. European Environment<br />

Agency: Luxembourg.<br />

(EPA) U.S. Environmental Protection Agency. 2004. Lower Mississippi<br />

Valley Ecosystem <strong>Restoration</strong> Initiative. [onl<strong>in</strong>e] URL:<br />

http://www.epa.gov/region4/programs/cbep/lowmiss.html<br />

(accessed 25 March 2004).<br />

Fleck L. 1979. Genesis and Development of a Scientifi c Fact.<br />

University of Chicago Press: Chicago.<br />

Funtowicz SO, Ravetz JR. 1995. Science for <strong>the</strong> post normal age.<br />

In Perspectives on Ecological Integrity, Westra L, Lemons<br />

J (eds). Kluwer Academic Publishers: Dordrecht, The<br />

Ne<strong>the</strong>rlands, 146–161.<br />

Gomez-Pompa A, Kaus A. 1992. Tam<strong>in</strong>g <strong>the</strong> wilderness myth.<br />

BioScience 42: 271–279.<br />

Gopal B. 2003. Aquatic biodiversity <strong>in</strong> arid and semi-arid zones<br />

of Asia and water management. In Conserv<strong>in</strong>g Biodiversity <strong>in</strong><br />

Arid Regions, Lemons J, Victor R, Schaffer D (Eds). Kluwer<br />

Academic Publishers: Norwell; 199–216.<br />

Higgs ES. 1997. What is good ecological restoration? Conservation<br />

Biology 11: 338–348.<br />

Hilborn R. 1997. Statistical hypo<strong>the</strong>sis test<strong>in</strong>g and decision<br />

<strong>the</strong>ory <strong>in</strong> fi sheries science. Fisheries 22: 19–20.<br />

Hull D. 1974. Philosophy of Biological Science. Prentice–Hall:<br />

Englewood Cliffs.<br />

Kaiser M, Storvik H (Eds). 2003. The Precautionary Pr<strong>in</strong>ciple:<br />

Between Research and Politics. The National Research<br />

Ethical Committee for Natural Science and Technology:<br />

Oslo, Norway.<br />

Karr JR, Chu EW. 1999. Restor<strong>in</strong>g Life <strong>in</strong> Runn<strong>in</strong>g Waters. Better<br />

Biological Monitor<strong>in</strong>g. Island Press: Wash<strong>in</strong>gton, DC.<br />

Kriebel D et al. 2001. The precautionary pr<strong>in</strong>ciple <strong>in</strong> environmental<br />

science. Environmental Health Perspectives 109: 871–876.<br />

Kuhn T. 1962. The Theory of Scientifi c Revolutions. University of<br />

Chicago Press: Chicago.<br />

Latour B. 1988. The politics of explanation: An alternative. In<br />

Knowledge and Refl exivity: New Frontiers <strong>in</strong> <strong>the</strong> Sociology<br />

of Knowledge, Woolgar S (Ed). Sage Publications: London;<br />

155–176.<br />

Lemons J (Ed). 1996. Scientifi c <strong>Uncerta<strong>in</strong>ty</strong> and Environmental<br />

Problem Solv<strong>in</strong>g. Blackwell Science: Cambridge.<br />

Lemons J, Brown D (Eds). 1995. Susta<strong>in</strong>able Development:<br />

Science, Ethics, and Public Policy. Kluwer Academic<br />

Publishers: Dordrecht, The Ne<strong>the</strong>rlands.<br />

Lemons J, Shrader-Frechette KS, Cranor C. 1997. The precautionary<br />

pr<strong>in</strong>ciple: Scientifi c uncerta<strong>in</strong>ty and type I and type II<br />

errors. Foundations of Science 2: 207–236.<br />

Light A, Higgs ES. 1996. The politics of ecological restoration.<br />

Environmental Ethics 18: 227–247.<br />

Lubchenco J et al. 1991. The susta<strong>in</strong>able biosphere <strong>in</strong>itiative: An<br />

ecological research agenda. Ecology 72: 371–412.<br />

MacMahon JA, Holl KD. 2001. Ecological restoration. A key to<br />

conservation biology’s future. In Conservation Biology.<br />

Research Priorities for <strong>the</strong> Next Decade, Soule ME, Orians GH<br />

(Eds). Island Press: Wash<strong>in</strong>gton, DC; 245–269.<br />

Mayo DG, Hollander RD (Eds). 1991. Acceptable Evidence.<br />

Oxford University Press: Oxford.<br />

(NRC) National Research Council. 1999. New Strategies for<br />

America’s Watersheds. National Academy Press, Wash<strong>in</strong>gton,<br />

DC.<br />

(NRC) National Research Council. 2004. Endangered and<br />

Threatened Fishes <strong>in</strong> <strong>the</strong> Klamath <strong>River</strong> Bas<strong>in</strong>: Causes of<br />

Decl<strong>in</strong>e and Strategies for Recovery. Committee on Endangered<br />

and Threatened Species <strong>in</strong> <strong>the</strong> Klamath <strong>River</strong> Bas<strong>in</strong>,<br />

Wash<strong>in</strong>gton, DC.<br />

Nienhuis PH, Leuven RSEW. 2001. <strong>River</strong> restoration and fl ood<br />

protection: Controversy or synergism? Hydrobiologia 444:<br />

85–99.<br />

Obeng LE. 1981. Man’s Impact on Tropical <strong>River</strong>s. In Perspectives<br />

<strong>in</strong> Runn<strong>in</strong>g Water Ecology, Lock MA, Williams DD (Eds).<br />

Plenum Press: New York; 265–288.<br />

Ogbeibu AE, Victor R. 1989. The effects of road and bridge<br />

construction on <strong>the</strong> bank–root macrobenthic <strong>in</strong>vertebrates<br />

of a sou<strong>the</strong>rn Nigerian stream. Environmental Pollution 56:<br />

85–100.


Peters RH. 1991. A Critique for Ecology. Cambridge University<br />

Press: Cambridge.<br />

Regier HA. 1995. Ecosystem <strong>in</strong>tegrity <strong>in</strong> a context of ecostudies<br />

as related to <strong>the</strong> Great Lakes Region. In Perspectives on Ecological<br />

Integrity, Westra L, Lemons J (Eds). Kluwer Academic<br />

Publishers: Dordrecht, The Ne<strong>the</strong>rlands, 88–101.<br />

Rolston H (III). 1988. Environmental Ethics, Duties to and Values<br />

<strong>in</strong> <strong>the</strong> Natural World. Temple University Press: Philadelphia.<br />

Rorty R. 1979. Philosophy and <strong>the</strong> Mirror of Nature. Pr<strong>in</strong>ceton<br />

University Press: New Jersey.<br />

Rzoska J. 1976. The Nile: Biology of an Ancient <strong>River</strong>. DW Junk<br />

Publishers: The Hague, The Ne<strong>the</strong>rlands.<br />

(SER) Society for Ecological <strong>Restoration</strong> Science and Policy<br />

Work<strong>in</strong>g Group. 2002. The SER Primer on Ecological<br />

<strong>Restoration</strong>. [onl<strong>in</strong>e] URL: www.ser.org (accessed 25 August<br />

2004).<br />

Service RF. 2003. NRC backs ecosystem-wide changes to save<br />

Klamath fi sh. Science 302: 765.<br />

Shrader-Frechette KS. 1982. Environmental impact assessment<br />

and <strong>the</strong> fallacy of unfi nished bus<strong>in</strong>ess. Environmental Ethics<br />

4: 37–48.<br />

Shrader-Frechette KS. 1994. Ethics of Scientifi c Research. Rowan<br />

& Littlefi eld Publisher, Inc.: Lanham.<br />

Shrader-Frechette KS. 1996. Methodological rules for four<br />

classes of scientifi c uncerta<strong>in</strong>ty. In Scientifi c <strong>Uncerta<strong>in</strong>ty</strong> and<br />

Environmental Problem Solv<strong>in</strong>g, Lemons J (Ed). Blackwell<br />

Science: Cambridge; 12–39.<br />

Shrader-Frechette KS, McCoy E. 1993. Method <strong>in</strong> Ecology.<br />

Cambridge University Press: Cambridge.<br />

Sirageld<strong>in</strong> I. 2002. Human prospects <strong>in</strong> an age of uncerta<strong>in</strong>ty and<br />

decl<strong>in</strong>e of rationality. Newsletter of <strong>the</strong> Economic Research<br />

Forum for <strong>the</strong> Arab Countries, Iran and Turkey 2: 9–11.<br />

Smits AJ, Niehuis PH, Leuven RSEW. 2000. New Approaches<br />

to <strong>River</strong> Management. Backhuys Publishers: Leiden, The<br />

Ne<strong>the</strong>rlands.<br />

Stewart-Oaten A. 1996. Goals <strong>in</strong> environmental monitor<strong>in</strong>g. In<br />

Detect<strong>in</strong>g Ecological Impacts: Concepts and Applications <strong>in</strong><br />

Coastal Habitats, Schmitt RJ, Osenbert CW (Eds). Academic<br />

Press: San Diego; 17–28.<br />

Sunste<strong>in</strong> C. 2002. Risk and Reason. Cambridge University Press.<br />

Cambridge.<br />

Sweeney S. 2000. Different means, shared ends: Environmental<br />

restoration and restoration ecology. Biologica 38: 129–139.<br />

(SWS) Society of Wetland Scientists position paper on <strong>the</strong> defi nitions<br />

of wetlands restoration. 2000. Ecological Applications 6:<br />

84–93.<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 13<br />

Tickner J. (Ed). 2002. Environmental Science and Preventive<br />

Public Policy. Island Press: Wash<strong>in</strong>gton DC.<br />

Tickner JA. 2003. Precaution: Environmental Science and<br />

Preventive Public Policy. Island Press: Wash<strong>in</strong>gton, DC.<br />

Victor R. 1998. Fish community changes associated with a small<br />

African reservoir. International Review of Hydrobiologia 83:<br />

603–610.<br />

Victor R, Brown CA. 1990. The food and feed<strong>in</strong>g habits of two<br />

species of characid fi sh <strong>in</strong> a perturbed West Africa river.<br />

Journal of African Zoology 104: 97–108.<br />

Victor R, Dickson DT. 1985. Macrobenthic <strong>in</strong>vertebrates of a<br />

perturbed stream <strong>in</strong> Sou<strong>the</strong>rn Nigeria. Environmental Pollution<br />

Series A 38: 99–107.<br />

Victor R, Meye J. 1994. Fur<strong>the</strong>r studies on <strong>the</strong> fi sh communities<br />

of a perturbed stream <strong>in</strong> Sou<strong>the</strong>rn Nigeria. Journal of Tropical<br />

Ecology 10: 627–632.<br />

Victor R, Ogbeibu AE. 1985. Macrobenthic <strong>in</strong>vertebrates of<br />

a stream fl ow<strong>in</strong>g through farmlands <strong>in</strong> Sou<strong>the</strong>rn Nigeria.<br />

Environmental Pollution Series A 39: 337–349.<br />

Victor R., Ogbeibu AE. 1986. Reconlonization of macrobenthic<br />

<strong>in</strong>vertebrates <strong>in</strong> a Nigerian stream after pesticide treatment and<br />

associated disruption. Environmental Pollution Series A 41:<br />

125–137.<br />

Victor R, Ogbeibu AE. 1991. Macrobenthic <strong>in</strong>vertebrates <strong>in</strong> <strong>the</strong><br />

erosional biotope of a Nigerian river. Tropical Zoology 4:<br />

1–12.<br />

Victor R. Onomivbori O. 1996. The effects of urban perturbations<br />

on <strong>the</strong> benthic macro<strong>in</strong>vertebrates of a sou<strong>the</strong>rn Nigerian<br />

stream. In Perspectives <strong>in</strong> Tropical Limnology, Schiemer F,<br />

Boland T (Eds). SPB Academic Publish<strong>in</strong>g: Amsterdam, The<br />

Ne<strong>the</strong>rlands; 233–238.<br />

Victor R, Tetteh JO. 1988. Fish communities of a perturbed<br />

stream <strong>in</strong> Sou<strong>the</strong>rn Nigeria. Nigeria Journal Tropical Ecology<br />

4: 49–59.<br />

Ward JV, Stanford JA. 1979. The Ecology of Regulated Streams.<br />

Plenum Press: New York.<br />

Westra L. 1995. Ecosystem <strong>in</strong>tegrity and susta<strong>in</strong>ability: The foundational<br />

value of <strong>the</strong> wild. In Perspectives on Ecological Integrity,<br />

Westra L, Lemons J (Eds). Kluwer Academic Publishers:<br />

Dordrecht, The Ne<strong>the</strong>rlands; 12–33.<br />

Westra L, Lemons J (Eds). 1995. Perspectives on Ecological<br />

Integrity. Kluwer Academic Publishers: Dordrecht, The<br />

Ne<strong>the</strong>rlands.<br />

Yoccoz NG. 1991. Use, overuse and misuse of signifi cance tests<br />

<strong>in</strong> evolutionary biology and ecology. Bullet<strong>in</strong> Ecological<br />

Society America 71: 106–111.


<strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat Edited by Stephen Darby and David Sear<br />

© 2008 John Wiley & Sons, Ltd<br />

2<br />

Sources of <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>River</strong><br />

<strong>Restoration</strong> Research<br />

2.1 INTRODUCTION: GENERAL SOURCES<br />

OF UNCERTAINTY<br />

The practice of science <strong>in</strong> support of river restoration is<br />

subject to four primary sources of uncerta<strong>in</strong>ty (see Chapters<br />

1 and 3 for additional/alternative views) so signifi cant<br />

that <strong>the</strong>y may prevent <strong>the</strong> restoration from achiev<strong>in</strong>g its<br />

goals. Firstly, <strong>the</strong> underly<strong>in</strong>g <strong>the</strong>ory applied by <strong>in</strong>vestigators<br />

to particular problem cases is imperfect and conta<strong>in</strong>s<br />

substantial gaps <strong>in</strong> explanatory and predictive capability.<br />

Secondly, <strong>the</strong> research process itself is subject to a variety<br />

of operational problems that <strong>in</strong>troduce uncerta<strong>in</strong>ty to <strong>the</strong><br />

use of science. Thirdly, <strong>the</strong> communication of scientifi c<br />

results to decision makers is often fraught with ambiguity<br />

derived from <strong>the</strong> scientifi c sender as well as <strong>the</strong> policy<br />

receiver. Fourthly, <strong>the</strong> scientists <strong>the</strong>mselves are subject to<br />

bias that generates doubt <strong>in</strong> <strong>the</strong> outcome of generat<strong>in</strong>g<br />

scientifi c products and apply<strong>in</strong>g <strong>the</strong>m. In <strong>the</strong> follow<strong>in</strong>g<br />

sections <strong>the</strong> issues for each of <strong>the</strong>se sources of uncerta<strong>in</strong>ty<br />

is outl<strong>in</strong>ed.<br />

2.2 UNCERTAINTY IN THEORY<br />

All science <strong>in</strong> support of river restoration beg<strong>in</strong>s with<br />

<strong>the</strong>ory, because it is <strong>the</strong>ory that allows <strong>the</strong> <strong>in</strong>vestigator to<br />

identify what to measure and how to construct a conceptual<br />

model that connects <strong>the</strong> measurements toge<strong>the</strong>r.<br />

Investigators perceive only those aspects of <strong>the</strong> river and<br />

its operations that <strong>the</strong>ory allows <strong>the</strong>m to see. Practitioners<br />

of fl uvial geomorphology tend to revere exist<strong>in</strong>g <strong>the</strong>ory as<br />

a sacrosanct start<strong>in</strong>g po<strong>in</strong>t but, like all sciences, geomorphology<br />

is <strong>in</strong> a state of constant change and revision. The<br />

William L. Graf<br />

Department of Geography, University of South Carol<strong>in</strong>a, USA<br />

change is sometimes gradual, as with <strong>the</strong> development and<br />

application of fundamental hydraulics to expla<strong>in</strong> river<br />

behavior that evolved over a period of several decades<br />

(Chang, 1998; Simons and Sentürk, 1992). Sometimes <strong>the</strong><br />

change is abrupt, as was <strong>the</strong> case with <strong>the</strong> <strong>in</strong>troduction of<br />

<strong>the</strong> concepts surround<strong>in</strong>g hydraulic geometry that burst<br />

upon <strong>the</strong> fl uvial geomorphology scene, became widely<br />

accepted <strong>in</strong> less than a decade and cont<strong>in</strong>ued <strong>in</strong> common<br />

use for several decades (Leopold, 1994). The result of this<br />

constantly chang<strong>in</strong>g <strong>the</strong>ory is that <strong>the</strong> geomorphologist<br />

work<strong>in</strong>g <strong>in</strong> 2007 may perceive a very different system than<br />

one work<strong>in</strong>g just a few years before or later, even though<br />

<strong>the</strong> physical system <strong>in</strong> all cases would be <strong>the</strong> same. <strong>Uncerta<strong>in</strong>ty</strong>,<br />

<strong>the</strong>refore, is <strong>in</strong>cluded <strong>in</strong> <strong>the</strong> application of science<br />

<strong>in</strong> its broadest sense.<br />

Ano<strong>the</strong>r source of ambiguity <strong>in</strong> <strong>the</strong>ory for river restoration<br />

is <strong>the</strong> regional specifi city that is built <strong>in</strong>to much<br />

of fl uvial <strong>the</strong>ory. Much of what we <strong>the</strong>orize about<br />

s<strong>in</strong>gle-thread meander<strong>in</strong>g rivers comes from research<br />

experience <strong>in</strong> northwest Europe and eastern North America<br />

(Knighton, 1998), yet <strong>the</strong> global applicability of this work<br />

is largely untested. Most of <strong>the</strong> streams of northwest<br />

Europe and eastern North America that have been <strong>in</strong>tensively<br />

<strong>in</strong>vestigated are relatively small on a world-wide<br />

basis, and though some generalities certa<strong>in</strong>ly must apply<br />

<strong>in</strong> many locales, <strong>the</strong> details may differ. Until <strong>the</strong> late<br />

1990s, much of <strong>the</strong> <strong>the</strong>ory for dryland rivers came from<br />

experiences <strong>in</strong> <strong>the</strong> American Southwest (Graf, 1988), but<br />

more recent <strong>in</strong>vestigations <strong>in</strong> Australia by Gerald Nanson,<br />

Steven Tooth and o<strong>the</strong>rs, for example, have shown that <strong>the</strong><br />

American experience is not applicable <strong>in</strong> all drylands<br />

(Nanson and Knighton, 1996).


16 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

If it is true that we must <strong>the</strong>orize based on what we<br />

know best, it must also be true that we are still limited <strong>in</strong><br />

<strong>the</strong> range of our collective experience. As a result, when<br />

we apply exist<strong>in</strong>g <strong>the</strong>ory <strong>in</strong> new geographic sett<strong>in</strong>gs, <strong>the</strong>re<br />

is reasonable doubt about <strong>the</strong> applicability of that <strong>the</strong>ory,<br />

at least <strong>in</strong> its totality. Geomorphologists commonly recognize<br />

that it is unwise to extend statistical models beyond<br />

<strong>the</strong> numerical ranges of <strong>the</strong> data. It is equally risky to<br />

extend geomorphological models beyond <strong>the</strong> geographic<br />

ranges of <strong>the</strong>ir orig<strong>in</strong>. The extension of <strong>the</strong>ory also forces<br />

us to consider how much of <strong>the</strong> geomorphology and<br />

hydrology of a particular river is unique, regardless of its<br />

geographic location. Each reach (a few kilometers long)<br />

of a stream is likely to be unique but <strong>the</strong> overall operation<br />

and form of a river (hundreds of kilometers long) is likely<br />

to have many similarities with o<strong>the</strong>r systems of similar<br />

magnitude. At <strong>the</strong> more extensive end of this range of<br />

magnitudes, generalizations are possible, while at <strong>the</strong> local<br />

end of <strong>the</strong> range of magnitudes, uniqueness becomes more<br />

apparent.<br />

The <strong>in</strong>completeness of most fl uvial geomorphologic<br />

<strong>the</strong>ory is also a source of uncerta<strong>in</strong>ty. This <strong>in</strong>completeness<br />

is <strong>in</strong> part purely a function of <strong>the</strong> natural river system, for<br />

which <strong>in</strong>vestigators have n<strong>in</strong>e fundamental operat<strong>in</strong>g variables,<br />

but for which <strong>the</strong>re are only a very few connective<br />

ma<strong>the</strong>matical functions (Leopold, Wolman, and Miller,<br />

1964). But an equally important limitation of exist<strong>in</strong>g<br />

<strong>the</strong>ory is its lack of recognition of human effects. Throughout<br />

much of <strong>the</strong> twentieth century (with a few exceptions),<br />

geomorphology as a science pursued explanation for<br />

‘natural’ rivers and many <strong>in</strong>vestigators made a conscious<br />

effort to avoid <strong>the</strong> confound<strong>in</strong>g <strong>in</strong>fl uence of technological<br />

<strong>in</strong>fl uences. It has only been <strong>in</strong> <strong>the</strong> last twenty years that<br />

those human <strong>in</strong>fl uences, pervasive and signifi cant <strong>in</strong> many<br />

rivers of <strong>the</strong> world, have <strong>the</strong>mselves become <strong>the</strong> objects<br />

of study (Costa et al., 1995; Graf, 2001). By defi nition,<br />

rivers subject to restoration have undergone changes<br />

result<strong>in</strong>g from human management and technology, but<br />

exist<strong>in</strong>g <strong>the</strong>ory is remarkably weak with respect to <strong>the</strong>se<br />

issues.<br />

The Colorado <strong>River</strong> <strong>in</strong> <strong>the</strong> Grand Canyon <strong>in</strong> <strong>the</strong> USA<br />

provides an example of <strong>the</strong> issues related to uncerta<strong>in</strong>ty <strong>in</strong><br />

<strong>the</strong>ory. Glen Canyon Dam, several kilometers upstream<br />

from <strong>the</strong> Grand Canyon controls <strong>the</strong> fl ow of <strong>the</strong> river, and<br />

especially reduces annual fl ood peaks to less than half of<br />

<strong>the</strong>ir former magnitude. The dam also reduces <strong>the</strong> sediment<br />

supply to <strong>the</strong> downstream canyon by more than 80%.<br />

As a result, <strong>the</strong> river has eroded sandy beaches and bars<br />

that once were common <strong>in</strong> <strong>the</strong> canyon (National Research<br />

Council, 1996). <strong>River</strong> restoration for <strong>the</strong> canyon <strong>in</strong>cluded<br />

re<strong>in</strong>troduction of moderate fl oods to move <strong>the</strong> available<br />

sediment from <strong>the</strong> channel fl oor to elevated positions,<br />

restor<strong>in</strong>g <strong>the</strong>se ecological niches. Despite considerable<br />

research, <strong>the</strong>re were no established <strong>the</strong>ories to predict <strong>the</strong><br />

response of <strong>the</strong> river to <strong>the</strong> artifi cial fl oods and although<br />

<strong>the</strong>re have been several fl ood-simulat<strong>in</strong>g releases from <strong>the</strong><br />

dam, <strong>the</strong> restoration results are not yet apparent.<br />

2.3 UNCERTAINTY IN RESEARCH<br />

Research us<strong>in</strong>g admittedly limited <strong>the</strong>ory <strong>in</strong> support of<br />

restoration is subject to uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> specifi cation<br />

of variables, assumptions, sampl<strong>in</strong>g, measurements and<br />

test<strong>in</strong>g of hypo<strong>the</strong>ses. The specifi cation or defi nition of<br />

variables, for example, is much entangled <strong>in</strong> <strong>the</strong> vagaries<br />

of science, law and personal perception of <strong>the</strong> researcher.<br />

Channel width provides an example. Most geomorphologists<br />

would agree that channel width is <strong>the</strong> distance across<br />

<strong>the</strong> active channel from one bank to <strong>the</strong> o<strong>the</strong>r, but <strong>the</strong><br />

application of this seem<strong>in</strong>gly simple proposition is devilishly<br />

diffi cult <strong>in</strong> many rivers. How should semi-permanent<br />

islands be taken <strong>in</strong>to account? What about ephemeral<br />

bars? How should width be determ<strong>in</strong>ed <strong>in</strong> <strong>the</strong> common<br />

circumstance where multiple sets of banks have resulted<br />

from episodic <strong>in</strong>cision or simply variable fl ows, which is<br />

often <strong>the</strong> case <strong>in</strong> arid, semi-arid, arctic or alp<strong>in</strong>e regions.<br />

Many legal systems also defi ne <strong>the</strong> channel as be<strong>in</strong>g<br />

‘between <strong>the</strong> banks,’ but do not specify which banks to use<br />

for <strong>the</strong> description (Graf, 1988).<br />

All geomorphological research <strong>in</strong>cludes assumptions<br />

which form ano<strong>the</strong>r source of uncerta<strong>in</strong>ty. The geographic<br />

and ecological complexity of rivers and <strong>the</strong>ir environments<br />

imply that when conduct<strong>in</strong>g <strong>in</strong>vestigations it is<br />

essential to focus on a few components and assume away<br />

<strong>the</strong> importance of variability <strong>in</strong> o<strong>the</strong>r factors that go<br />

unmeasured. In geomorphology, hydrology and eng<strong>in</strong>eer<strong>in</strong>g<br />

studies that support restoration, <strong>in</strong>vestigators often<br />

assume stationarity of <strong>the</strong> hydro–climatic processes rul<strong>in</strong>g<br />

<strong>the</strong> river. Stationarity means <strong>in</strong>vestigators assume that<br />

<strong>the</strong> underly<strong>in</strong>g statistical distributions describ<strong>in</strong>g climatic<br />

variables important to river processes are unchang<strong>in</strong>g.<br />

Standard magnitude/frequency analysis <strong>in</strong>cludes this<br />

assumption so that <strong>the</strong> researcher can address o<strong>the</strong>r variables<br />

of <strong>in</strong>terest to planners, <strong>in</strong>clud<strong>in</strong>g <strong>the</strong> return <strong>in</strong>tervals<br />

for various magnitudes of discharge. However, climate is<br />

anyth<strong>in</strong>g but stationary, and its variation is highly likely<br />

to <strong>in</strong>fl uence <strong>the</strong> statistical distributions upon which return<br />

<strong>in</strong>terval concepts depend. This variation is also likely to<br />

be signifi cant to <strong>the</strong> fl uvial system over time scales as<br />

short as decades, scales that encompass <strong>the</strong> likely project<br />

life of most restoration efforts. Predictions for <strong>the</strong> nearterm<br />

future of a few decades are <strong>the</strong>refore uncerta<strong>in</strong><br />

because <strong>the</strong> effects of expected climatic changes are not<br />

part of <strong>the</strong> analysis (see Chapter 13).


Sampl<strong>in</strong>g processes <strong>in</strong> fl uvial geomorphology also cast<br />

doubt on <strong>the</strong> confi dence users may have <strong>in</strong> <strong>the</strong> reliability<br />

of <strong>the</strong> result<strong>in</strong>g data. <strong>River</strong>s present <strong>the</strong> researcher with<br />

long l<strong>in</strong>es or corridors, and if <strong>the</strong> <strong>in</strong>vestigator uses cross<br />

sections or po<strong>in</strong>t samples, <strong>the</strong> selection of <strong>the</strong> locations of<br />

<strong>the</strong>se sample po<strong>in</strong>ts may strongly <strong>in</strong>fl uence <strong>the</strong> result<strong>in</strong>g<br />

data. The spac<strong>in</strong>g of meanders, riffl es and pools <strong>in</strong>troduces<br />

some regularity to <strong>the</strong> spac<strong>in</strong>g of important geomorphic<br />

and hydraulic features of rivers, so that if <strong>in</strong>vestigators use<br />

regular spac<strong>in</strong>g for sample cross sections or sites, that<br />

spac<strong>in</strong>g may co<strong>in</strong>cide with <strong>the</strong> spac<strong>in</strong>g of particular<br />

characteristics of <strong>the</strong> channel. For example, sample sites<br />

might occur only on riffl es, or only <strong>in</strong> pools, giv<strong>in</strong>g a false<br />

picture of <strong>the</strong> system. Truly random spac<strong>in</strong>g for samples<br />

may be statistically desirable, but a realistic view of geomorphic<br />

and hydrologic conditions demands that all <strong>the</strong><br />

various environments of <strong>the</strong> channel be <strong>in</strong>cluded <strong>in</strong> <strong>the</strong><br />

sample, someth<strong>in</strong>g that is not possible without some<br />

understand<strong>in</strong>g of <strong>the</strong> basic spatial framework of <strong>the</strong><br />

system.<br />

<strong>Uncerta<strong>in</strong>ty</strong> may be lessened if scale is part of <strong>the</strong> plann<strong>in</strong>g<br />

process for select<strong>in</strong>g sample sites. Sample schemes<br />

couched <strong>in</strong> <strong>the</strong> concept of river reaches permit bracket<strong>in</strong>g<br />

of relevant parts of <strong>the</strong> stream. A river reach is from one<br />

to a few kilometers <strong>in</strong> length and conta<strong>in</strong>s similar geomorphic<br />

conditions throughout its extent, sometimes with<br />

repetitive and alternat<strong>in</strong>g channel confi gurations such as<br />

pools and riffl es, pools and rapids, or meanders. Several<br />

reaches may make up a river segment. Geological boundaries,<br />

confl uences with major tributaries or human structures<br />

such as dams and diversion works create <strong>the</strong> upstream<br />

and downstream boundaries of each segment. Sampl<strong>in</strong>g<br />

schemes constructed with <strong>the</strong> reaches and segments as<br />

geographic frameworks are more likely to be <strong>in</strong>formative<br />

for restoration work than truly regular or truly random<br />

samples. Project design for restoration requires a clear<br />

assessment of <strong>the</strong> appropriate dimensions and spac<strong>in</strong>g of<br />

repetitions to <strong>in</strong>sure long term stability. If <strong>the</strong> dimensions<br />

and spac<strong>in</strong>g are not refl ective of <strong>the</strong> restored hydraulic<br />

conditions, <strong>the</strong> restored system will be unstable and may<br />

dis<strong>in</strong>tegrate.<br />

The example of restoration of <strong>the</strong> Platte <strong>River</strong> <strong>in</strong><br />

Nebraska <strong>in</strong> <strong>the</strong> USA illustrates <strong>the</strong> role of uncerta<strong>in</strong>ty <strong>in</strong><br />

research. Upstream fl ood control dams have brought about<br />

great changes <strong>in</strong> <strong>the</strong> hydrologic regime and geomorphology<br />

of <strong>the</strong> Platte, a serious issue because <strong>the</strong> river<br />

hosts several endangered bird species. The river orig<strong>in</strong>ally<br />

<strong>in</strong>cluded valuable habitats for birds, particularly <strong>the</strong><br />

whoop<strong>in</strong>g crane, but <strong>the</strong> hydrologic and geomorphic<br />

changes have reduced <strong>the</strong>ir habitat. <strong>River</strong> restoration<br />

<strong>in</strong>cludes return<strong>in</strong>g <strong>the</strong> river to conditions closer to those<br />

that prevailed before <strong>the</strong> dams were <strong>in</strong> place. Features such<br />

Sources of <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> Research 17<br />

as multiple channels, high and low islands, and complex<br />

bars are essential to <strong>the</strong> restoration. Research to support<br />

<strong>the</strong> restoration has been under way for more than a decade<br />

but has produced results that are diffi cult to <strong>in</strong>terpret<br />

(National Research Council, 2004). Cross-sectional<br />

surveys, for example, are numerous but not often conducted<br />

at <strong>the</strong> same places through time so that sampl<strong>in</strong>g<br />

is an issue. The fl ow parameters that are most important<br />

to <strong>the</strong> species may not be <strong>the</strong> same parameters that are<br />

important to <strong>the</strong> geomorphology. Stationarity is a particular<br />

problem <strong>in</strong> deal<strong>in</strong>g with <strong>the</strong> hydrologic records of <strong>the</strong><br />

Platte because <strong>the</strong> river is located on <strong>the</strong> boundary between<br />

sub-humid and sub-arid regions and is subject to climatic<br />

fl uctuations on decadal and century-long periods which<br />

casts doubts on shorter term records.<br />

2.4 UNCERTAINTY IN COMMUNICATION<br />

The successful use of science <strong>in</strong> formulat<strong>in</strong>g public policy<br />

for <strong>the</strong> restoration of rivers relies on accurate communication<br />

between researchers and decision makers, but this<br />

connection sometimes suffers from fail<strong>in</strong>gs by both participants.<br />

Scientists are usually accustomed to communicat<strong>in</strong>g<br />

with each o<strong>the</strong>r us<strong>in</strong>g a specifi c shared language of<br />

technical terms. Even terms shared with <strong>the</strong> general language<br />

may take on specifi c shad<strong>in</strong>gs of mean<strong>in</strong>g when<br />

used by specialists communicat<strong>in</strong>g with each o<strong>the</strong>r. A<br />

steep river gradient, for example, calls to m<strong>in</strong>d a general<br />

picture for most geomorphologists, perhaps of a step-pool<br />

sequence for mounta<strong>in</strong> streams or a braided channel <strong>in</strong><br />

o<strong>the</strong>r sett<strong>in</strong>gs. For <strong>the</strong> decision maker, a steep river may<br />

be one with water falls. More importantly, specialized<br />

terms and terms with nuances do not work well when <strong>the</strong><br />

listener or reader is not tra<strong>in</strong>ed or experienced <strong>in</strong> <strong>the</strong> scientifi<br />

c specialty. As a result, scientists have an obligation<br />

to use language without jargon and without assumptions<br />

when communicat<strong>in</strong>g with decision makers and <strong>the</strong><br />

<strong>in</strong>formed public, a task that seem<strong>in</strong>gly challenges many<br />

specialists.<br />

Decision makers, on <strong>the</strong> o<strong>the</strong>r hand, have poorly<br />

<strong>in</strong>formed expectations regard<strong>in</strong>g <strong>the</strong>ir scientifi c advisors.<br />

They expect clear, unambiguous answers to <strong>the</strong>ir questions<br />

and specifi c, robust predictions of future processes that<br />

might result from a variety of potential decisions. Science,<br />

however, rarely offers truly unambiguous conclusions, and<br />

caveats are many <strong>in</strong> applied geomorphology. Often, it is<br />

possible to predict <strong>the</strong> direction of change, but <strong>the</strong>re is<br />

greater trouble predict<strong>in</strong>g <strong>the</strong> magnitude of that change.<br />

To a public servant attempt<strong>in</strong>g to protect property values,<br />

such predictions may lack <strong>the</strong> required level of comfort.<br />

A common error <strong>in</strong> communication that results <strong>in</strong><br />

uncerta<strong>in</strong>ty occurs when <strong>the</strong> decision maker requests


18 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

specifi c numbers, such as ‘what will be <strong>the</strong> stable, long<br />

term width of <strong>the</strong> restored channel’, a reasonable question<br />

<strong>in</strong> <strong>the</strong> public policy arena. However, most scientifi c and<br />

eng<strong>in</strong>eer<strong>in</strong>g models predict some most likely value, enveloped<br />

<strong>in</strong> error bars. The most effective (and honest)<br />

report <strong>in</strong>cludes <strong>the</strong> most probable value but also <strong>in</strong>cludes<br />

some discussion of potential deviation from that expected<br />

value. Report<strong>in</strong>g of potential errors and ranges of possible<br />

values ra<strong>the</strong>r than s<strong>in</strong>gle numbers protects <strong>the</strong> <strong>in</strong>terest of<br />

everyone <strong>in</strong>volved and produces more reasonable expectations<br />

on <strong>the</strong> part of <strong>the</strong> general public.<br />

<strong>Restoration</strong> of <strong>the</strong> Elwha <strong>River</strong> <strong>in</strong> <strong>the</strong> state of Wash<strong>in</strong>gton<br />

exemplifi es <strong>the</strong> issues surround<strong>in</strong>g uncerta<strong>in</strong>ty <strong>in</strong> communication.<br />

National legislators directed that <strong>the</strong> river be<br />

restored to its ‘natural’ condition for <strong>the</strong> benefi t of endangered<br />

salmon, which use <strong>the</strong> river for spawn<strong>in</strong>g dur<strong>in</strong>g<br />

annual migrations (US House of Representatives, 1992).<br />

The method of restoration centers on <strong>the</strong> removal of two<br />

large dams. However, from a scientifi c standpo<strong>in</strong>t, it is<br />

impossible to meet <strong>the</strong> requirements of <strong>the</strong> law, because<br />

even with <strong>the</strong> removal of <strong>the</strong> dams, <strong>the</strong> river will not even<br />

approach truly natural conditions. Upstream land use,<br />

<strong>in</strong>clud<strong>in</strong>g logg<strong>in</strong>g, have altered <strong>the</strong> basic hydrologic and<br />

sediment regimes of <strong>the</strong> river, and <strong>in</strong> downstream areas<br />

levees and o<strong>the</strong>r structures prevent natural river processes.<br />

In fact, ‘natural’ conditions are a model to which restoration<br />

might aspire, but <strong>the</strong> pr<strong>in</strong>ciple as a true objective is<br />

unworkable.<br />

2.5 BIAS<br />

A fi nal source of uncerta<strong>in</strong>ty <strong>in</strong> science for river restoration<br />

lies with<strong>in</strong> <strong>the</strong> <strong>in</strong>tellect of <strong>the</strong> researchers <strong>the</strong>mselves.<br />

All geomorphologists are products of <strong>the</strong>ir cultural backgrounds,<br />

academic tra<strong>in</strong><strong>in</strong>g, experiences and personal<br />

characteristics (see Chapters 1 and 4). Researchers raised<br />

<strong>in</strong> <strong>in</strong>tellectually liberal surround<strong>in</strong>gs may be more question<strong>in</strong>g<br />

of established <strong>the</strong>ories than those raised <strong>in</strong> more<br />

rule-bound sett<strong>in</strong>gs, and each of <strong>the</strong>se types is likely to<br />

approach problems, evidence and methods with different<br />

biases and preferences. Gender may play a subtle, but as<br />

yet relatively unexplored role <strong>in</strong> our approaches to science.<br />

Academic tra<strong>in</strong><strong>in</strong>g for fl uvial geomorphology is highly<br />

variable from one group of teachers to ano<strong>the</strong>r, with some<br />

groups emphasiz<strong>in</strong>g a stochastic ra<strong>the</strong>r than a determ<strong>in</strong>istic<br />

approach. O<strong>the</strong>r contrast<strong>in</strong>g styles <strong>in</strong>clude greater<br />

emphasis on geographic approaches as opposed to those<br />

more strongly oriented toward eng<strong>in</strong>eer<strong>in</strong>g, research<br />

designs that emphasize small-scale (particle-sized) as<br />

opposed to ecosystem (or landscape) scale perspectives,<br />

or empirical ra<strong>the</strong>r than model-based approaches. Once<br />

tra<strong>in</strong>ed, <strong>the</strong> researcher is fur<strong>the</strong>r formed by personal fi eld<br />

experiences, with extensive backgrounds <strong>in</strong> dryland sett<strong>in</strong>gs<br />

produc<strong>in</strong>g a different perspective than experiences<br />

dom<strong>in</strong>ated by humid subtropical environments, tropical<br />

sett<strong>in</strong>gs or polar landscapes. F<strong>in</strong>ally, scientists are just like<br />

everyone else: some are stubborn, some are open m<strong>in</strong>ded;<br />

some are methodical while <strong>the</strong>ir colleagues make successful<br />

leaps of logic; and some are quick to reach judgements<br />

while o<strong>the</strong>rs seem never to reach closure on <strong>the</strong>ir<br />

conclusions.<br />

There is noth<strong>in</strong>g <strong>in</strong>herently wrong with biases, for to be<br />

biased is to be human. In river restoration, however, and<br />

especially when deal<strong>in</strong>g with decision makers, <strong>the</strong> wise<br />

scientists fi nd ways to communicate <strong>the</strong>ir biases to <strong>the</strong><br />

consumers of <strong>the</strong> scientifi c products. If <strong>the</strong> consumer<br />

(policy maker or <strong>in</strong>formed citizen) recognizes <strong>the</strong> biases<br />

and takes <strong>the</strong>m <strong>in</strong>to account, <strong>the</strong> <strong>in</strong>terpretation and application<br />

of <strong>the</strong> scientifi c products is confi dent by all parties<br />

<strong>in</strong>volved. Admission of biases by <strong>the</strong> researcher, usually<br />

<strong>in</strong> <strong>the</strong> form of a brief disclosure that clearly defi nes<br />

<strong>the</strong> schools of thought and experience of <strong>the</strong> researcher,<br />

enhances <strong>the</strong> professionalism of <strong>the</strong> <strong>in</strong>vestigator and fairly<br />

discloses to <strong>the</strong> consumer <strong>the</strong> background of <strong>the</strong> knowledge<br />

that forms <strong>the</strong> basis of decisions perta<strong>in</strong><strong>in</strong>g to public<br />

resources. In unusual cases, critics may contest decisions<br />

<strong>in</strong> adm<strong>in</strong>istrative hear<strong>in</strong>gs or <strong>in</strong> court proceed<strong>in</strong>gs, two<br />

venues where biases are likely to be revealed and explored<br />

<strong>in</strong> a combative environment. If biases have previously<br />

been revealed, <strong>the</strong>y lack s<strong>in</strong>ister overtones <strong>in</strong> a strategy<br />

that benefi ts both researcher and consumer.<br />

2.6 CONCLUSION<br />

<strong>Uncerta<strong>in</strong>ty</strong> is a fact of life as well as an <strong>in</strong>escapable<br />

feature of scientifi c research and decision mak<strong>in</strong>g for river<br />

restoration. Therefore, <strong>the</strong> researcher has two essential<br />

options: ei<strong>the</strong>r ignore <strong>the</strong> uncerta<strong>in</strong>ty and hope that it is<br />

not debilitat<strong>in</strong>g for <strong>the</strong> project at hand, or accept <strong>the</strong> uncerta<strong>in</strong>ty<br />

and use it as a feature of <strong>the</strong> research. The researcher<br />

can <strong>in</strong>vestigate <strong>the</strong> uncerta<strong>in</strong>ty, quantify it <strong>in</strong> some cases,<br />

and reveal it <strong>in</strong> explicit terms when report<strong>in</strong>g results. In<br />

this latter approach, uncerta<strong>in</strong>ty becomes an <strong>in</strong>tegral part<br />

of research for river restoration, a feature of <strong>the</strong> work that<br />

is a welcome challenge to be embraced and used to achieve<br />

a more effective end product. Project design may <strong>in</strong>clude<br />

a variety of channel dimensions and characteristics, for<br />

example, and avoid rely<strong>in</strong>g on a s<strong>in</strong>gle rigidly defi ned<br />

morphology, so that if orig<strong>in</strong>al understand<strong>in</strong>gs of <strong>the</strong><br />

system are not exactly correct <strong>the</strong> fi nal project will have<br />

some fl exibility. In o<strong>the</strong>r cases, it may be wise to simply<br />

allot more space for channel changes <strong>in</strong> <strong>the</strong> designed<br />

project to accommodate unforeseen adjustments. By<br />

deal<strong>in</strong>g directly with uncerta<strong>in</strong>ty, researcher and decision


maker <strong>in</strong>crease <strong>the</strong> probability <strong>in</strong> successfully restor<strong>in</strong>g a<br />

river with enhanced environmental and social benefi ts.<br />

REFERENCES<br />

Chang HH. 1998. Fluvial Processes <strong>in</strong> <strong>River</strong> Eng<strong>in</strong>eer<strong>in</strong>g.<br />

Krieger: Malabar, Florida.<br />

Collier MP, Webb RH, Andrews ED. 1997. Experimental fl ood<strong>in</strong>g<br />

<strong>in</strong> <strong>the</strong> Grand Canyon. Scientifi c American 276: 82–89.<br />

Costa JE, Miller AJ, Potter KW, Wilcock PR (Eds). 1995. Natural<br />

and Anthropogenic Infl uences <strong>in</strong> Fluvial Geomorphology: The<br />

Wolman Volume, Geophysical Monograph 89, American Geophysical<br />

Union: Wash<strong>in</strong>gton, DC.<br />

Graf WL. 1988. Defi nition of fl ood pla<strong>in</strong>s along arid-region<br />

rivers. In: Baker VR, Kochel RC, Patton PC (Eds), Flood<br />

Geomorphology, John Wiley & Sons, Inc.: New York, NY;<br />

231–242.<br />

Graf WL. 2001. Damage Control: Dams and <strong>the</strong> physical <strong>in</strong>tegrity<br />

of America’s rivers. Annals of <strong>the</strong> Association of American<br />

Geographers 91: 1–27.<br />

Sources of <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> Research 19<br />

Knighton D. 1998. Fluvial Forms and Processes: A New Perspective.<br />

Arnold: London.<br />

Leopold LB, Wolman MG, Miller JP. 1964. Fluvial Processes <strong>in</strong><br />

Geomorphology. WH Freeman: San Francisco, California.<br />

Leopold LB. 1994. A View of <strong>the</strong> <strong>River</strong>. Harvard University Press:<br />

Cambridge, Massachusetts.<br />

Nanson GC, Knighton AD. 1996. Anabranch<strong>in</strong>g rivers: Their<br />

cause, character and classifi cation. Earth Surface Processes<br />

and Landforms 21: 217–239.<br />

National Research Council. 1996. <strong>River</strong> Resource Management<br />

<strong>in</strong> <strong>the</strong> Grand Canyon. National Academy Press: Wash<strong>in</strong>gton,<br />

DC.<br />

National Research Council. 2004. Endangered and Threatened<br />

Species of <strong>the</strong> Platte <strong>River</strong>. National Academy Press: Wash<strong>in</strong>gton,<br />

D.C.<br />

Simons DB, Sentürk F. 1992. Sediment Transport Technology.<br />

Water Resources Publications: Littleton, Colorado.<br />

US House of Representatives. 1992. Jo<strong>in</strong>t Hear<strong>in</strong>gs on HR 4844,<br />

Elwha <strong>River</strong> Ecosystem and Fisheries <strong>Restoration</strong> Act, 102nd<br />

Congress, 2nd Session. Government Pr<strong>in</strong>t<strong>in</strong>g Offi ce: Wash<strong>in</strong>gton,<br />

DC.


3.1 INTRODUCTION<br />

<strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat Edited by Stephen Darby and David Sear<br />

© 2008 John Wiley & Sons, Ltd<br />

3<br />

The Scope of Uncerta<strong>in</strong>ties <strong>in</strong><br />

<strong>River</strong> <strong>Restoration</strong><br />

Joseph M. Wheaton 1 , Stephen E. Darby 2 and David A. Sear 2<br />

1 Institute of Geography and Earth Sciences, The University of Wales, UK<br />

2 School of Geography, University of Southampton, UK<br />

The science and practice of river restoration are both<br />

still very much <strong>in</strong> <strong>the</strong>ir adolescence (Palmer et al., 1997).<br />

Yet, both have been graced with fund<strong>in</strong>g and support<br />

from a diverse range of <strong>in</strong>terest groups (Malakoff, 2004).<br />

One of <strong>the</strong> premises of this book is that if fund<strong>in</strong>g is to<br />

cont<strong>in</strong>ue to be allocated to river restoration, it will have to<br />

be shown that river restoration is ‘work<strong>in</strong>g’ (see Preface;<br />

Wissmar and Bisson, 2003c). Defi nitions of ‘work<strong>in</strong>g’<br />

(often equated with success) are understandably subjective<br />

and vulnerable to uncerta<strong>in</strong>ties <strong>in</strong> <strong>the</strong> river restoration<br />

process, societal values, <strong>the</strong> fl uvial system and ecosystem<br />

response to restoration management activities. Davis and<br />

Slobodk<strong>in</strong> (2004) argued that defi n<strong>in</strong>g restoration goals<br />

and objectives is rightfully a value-based activity, as<br />

opposed to scientifi c activity. Each activity is <strong>in</strong>herently<br />

uncerta<strong>in</strong>. Paradoxically, <strong>the</strong> uncerta<strong>in</strong>ties <strong>in</strong>fl uenc<strong>in</strong>g<br />

river restoration projects are rarely recognised or quantifi<br />

ed, much less reported to stakeholders or <strong>the</strong> public<br />

(Walters, 1997).<br />

The topic of uncerta<strong>in</strong>ty <strong>in</strong> river restoration is riddled<br />

with complexity and confusion. Indeed, uncerta<strong>in</strong>ty manifests<br />

itself <strong>in</strong> many ways, as established <strong>in</strong> Chapters 1 and<br />

2. Lemons and Victor (see Chapter 1) have already illustrated<br />

how deep <strong>the</strong> value-laden dimensions of uncerta<strong>in</strong>ty<br />

lies, not just <strong>in</strong> decision mak<strong>in</strong>g, but <strong>in</strong> scientifi c research<br />

as well. Graf (see Chapter 2) expanded on this <strong>the</strong>me,<br />

cit<strong>in</strong>g uncerta<strong>in</strong>ties from <strong>the</strong>ories, <strong>the</strong> research itself, communication<br />

and biases among <strong>in</strong>vestigators. He concluded<br />

that <strong>the</strong> research can ei<strong>the</strong>r ‘ignore <strong>the</strong> uncerta<strong>in</strong>ty and<br />

hope that it is not debilitat<strong>in</strong>g for <strong>the</strong> project at hand,<br />

or accept <strong>the</strong> uncerta<strong>in</strong>ty and use it as a feature of <strong>the</strong><br />

research.’ In this chapter <strong>the</strong> rich topic of uncerta<strong>in</strong>ty is<br />

presented <strong>in</strong> a broader, more generic context. This foundation<br />

is <strong>in</strong>tended to help separate <strong>the</strong> sources and types of<br />

uncerta<strong>in</strong>ties that <strong>the</strong> various authors <strong>in</strong> this book present,<br />

and meanwhile unravel some of <strong>the</strong> ambiguities surround<strong>in</strong>g<br />

uncerta<strong>in</strong>ty <strong>in</strong> river restoration.<br />

As cautioned earlier, potentially signifi cant uncerta<strong>in</strong>ties<br />

are rarely recognised, much less explicitly dealt<br />

with <strong>in</strong> river restoration. Hence, a lexicon and typology<br />

for uncerta<strong>in</strong>ty is outl<strong>in</strong>ed fi rstly <strong>in</strong> this chapter. This is<br />

done to dispel <strong>the</strong> notion of a certa<strong>in</strong> world with certa<strong>in</strong><br />

outcomes with<strong>in</strong> <strong>the</strong> broader scope of types and sources<br />

of uncerta<strong>in</strong>ty. A return specifi cally to river restoration<br />

<strong>the</strong>n follows to identify types of uncerta<strong>in</strong>ties us<strong>in</strong>g <strong>the</strong><br />

above-mentioned typology. The tremendous diversity of<br />

river restoration <strong>in</strong> <strong>the</strong> context of uncerta<strong>in</strong>ties aris<strong>in</strong>g<br />

from restoration motives, notions and approaches are considered.<br />

A case will be made that a basic strategy for<br />

deal<strong>in</strong>g with uncerta<strong>in</strong>ty is needed by <strong>the</strong> river restoration<br />

community to allow both <strong>the</strong> community and <strong>in</strong>dividual<br />

<strong>in</strong>vestigators or practitioners to:<br />

explore <strong>the</strong> potential signifi cance (both <strong>in</strong> terms of<br />

unforeseen consequences and welcome surprises) or<br />

<strong>in</strong>signifi cance of uncerta<strong>in</strong>ties;<br />

effectively communicate uncerta<strong>in</strong>ties;<br />

eventually make adaptive, but transparent, decisions <strong>in</strong><br />

<strong>the</strong> face of uncerta<strong>in</strong>ty.


22 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

F<strong>in</strong>ally, it will be argued that amongst <strong>the</strong> various strategies<br />

for deal<strong>in</strong>g with uncerta<strong>in</strong>ty, <strong>the</strong> only strategy that might<br />

provide <strong>the</strong>se aims is one of embrac<strong>in</strong>g uncerta<strong>in</strong>ty.<br />

3.1.1 The Status Quo <strong>in</strong> <strong>River</strong> <strong>Restoration</strong><br />

The rapid rise and <strong>in</strong>ternational popularity of river restoration<br />

is both encourag<strong>in</strong>g and worrisome (Kondolf, 1996).<br />

Although sparse examples dat<strong>in</strong>g back to <strong>the</strong> 1930s exist1 ,<br />

river restoration has primarily developed on <strong>the</strong> coat tails<br />

of <strong>the</strong> environmental awareness movement of <strong>the</strong> late<br />

1970s (Graf, 1996; Sear, 1994). It is encourag<strong>in</strong>g that so<br />

much enthusiasm exists to restore rivers. Yet, it is <strong>in</strong>terest<strong>in</strong>g<br />

to note <strong>the</strong> societal choices between some mix of<br />

reactive restoration efforts <strong>in</strong> response to damage already<br />

done, as opposed to pro-active conservation actions to<br />

prevent fur<strong>the</strong>r damage (Boon, 1998). The <strong>in</strong>ternational<br />

popularity of river restoration is evident <strong>in</strong> <strong>the</strong> restoration<br />

literature (e.g. restoration <strong>in</strong> 21 different countries reported<br />

<strong>in</strong> Nijland and Cals, 2000), restoration databases (e.g. <strong>the</strong><br />

United K<strong>in</strong>gdom <strong>River</strong> <strong>Restoration</strong> Centre (RRC), United<br />

States Environmental Protection Agency (EPA)) 2 and an<br />

International <strong>River</strong> <strong>Restoration</strong> Survey3 launched by<br />

Wheaton et al. (2004c) with respondents from 34 different<br />

countries. In Denmark alone, 1068 restoration projects had<br />

been completed by Danish regional authorities by 1998<br />

(Hansen and Iversen, 1998); whereas <strong>in</strong> <strong>the</strong> United States,<br />

Malakoff (2004) reported that by 2004 more than $US10<br />

billion had been spent on a total of more than 30 000<br />

projects. The popularity of river restoration is apparent <strong>in</strong><br />

<strong>in</strong>ternational, national, regional and local public policy that<br />

actively promotes, requires and, <strong>in</strong> some cases, funds river<br />

restoration efforts (Jungwirth et al., 2002). However, <strong>the</strong>ir<br />

effectiveness is constra<strong>in</strong>ed by limited funds and scope to<br />

deal with closely related land use issues and o<strong>the</strong>r sociopolitical<br />

goals (Tockner and Stanford, 2002).<br />

Despite <strong>the</strong> popularity of river restoration <strong>in</strong> <strong>the</strong> developed<br />

nations of <strong>the</strong> world, <strong>the</strong> global decl<strong>in</strong>e of <strong>the</strong> physical<br />

and ecological <strong>in</strong>tegrity of rivers is diffi cult to overstate<br />

(Jungwirth et al., 2002; Vitousek et al., 1997). Indeed,<br />

1 The United States Department of Agriculture Forest Service<br />

started undertak<strong>in</strong>g ‘stream improvement’ <strong>in</strong> <strong>the</strong> 1930s with <strong>the</strong><br />

<strong>in</strong>tent of <strong>in</strong>creas<strong>in</strong>g salmonid production (Everest and Sedell,<br />

1984).<br />

2 RRC Database <strong>in</strong>cludes over 750 projects with<strong>in</strong> <strong>the</strong> United<br />

K<strong>in</strong>gdom: http://www.<strong>the</strong>rrc.co.uk; <strong>the</strong> USEPA <strong>River</strong> Corridor<br />

and Wetland <strong>Restoration</strong> Database <strong>in</strong>cludes over 600 projects<br />

throughout <strong>the</strong> United States: http://yosemite.epa.gov/water/<br />

restorat.nsf/rpd-2a.htm.<br />

3 Complete real time results, background <strong>in</strong>formation and forthcom<strong>in</strong>g<br />

<strong>in</strong>terpretations are available on <strong>the</strong> web: http://www.geog.<br />

soton.ac.uk/users/WheatonJ/<strong>Restoration</strong>Survey_Cover.asp.<br />

most restoration efforts still pale <strong>in</strong>to signifi cance relative<br />

to expand<strong>in</strong>g anthropogenic impacts on river<strong>in</strong>e landscapes<br />

(Tockner and Stanford, 2002). Even <strong>in</strong> parts of <strong>the</strong><br />

world where numerous river restoration efforts are already<br />

underway (i.e. Europe, North America and Australia), wetlands<br />

are actively be<strong>in</strong>g dra<strong>in</strong>ed and fi lled, rivers are still<br />

diverted and regulated, urban growth is encroach<strong>in</strong>g <strong>in</strong>to<br />

fl oodpla<strong>in</strong>s and headwaters, while we cont<strong>in</strong>ue to permanently<br />

alter bas<strong>in</strong> hydrology and fragment habitats (Coll<strong>in</strong>s<br />

et al., 2000; Moss, 2004; Mount, 1995). These problems<br />

pose even larger threats <strong>in</strong> <strong>the</strong> develop<strong>in</strong>g nations of<br />

<strong>the</strong> world (Marmulla, 2001). Over 250 new major dams<br />

become operational worldwide annually and 75 are<br />

planned for <strong>the</strong> Amazon Bas<strong>in</strong> alone (Rob<strong>in</strong>son et al.,<br />

2002). It seems logical that preservation should be easier<br />

to achieve than restoration (Frissell et al., 1993), but <strong>the</strong>re<br />

seems to be excessive confi dence <strong>in</strong> <strong>the</strong> ability to restore<br />

(Stewardson and Ru<strong>the</strong>rfurd, see Chapter 5), sometimes<br />

reduc<strong>in</strong>g restoration to a mitigation measure justify<strong>in</strong>g<br />

planned impacts or ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g <strong>the</strong> status quo. Both conservation<br />

and restoration are based on <strong>the</strong> transformation<br />

of uncerta<strong>in</strong> science and uncerta<strong>in</strong> notions of what is<br />

natural, ecosystem <strong>in</strong>tegrity and physical <strong>in</strong>tegrity <strong>in</strong>to<br />

societal goals (Graf, 2001; Lemons and Victor, see Chapter<br />

1). Additionally, <strong>the</strong> good <strong>in</strong>tentions of restoration projects<br />

may lead to un<strong>in</strong>tended but often foreseeable consequences.<br />

Even if society is will<strong>in</strong>g to make diffi cult sociopolitical<br />

decisions to support preservation and restoration<br />

of rivers, <strong>the</strong>re is no guarantee of desired outcomes<br />

follow<strong>in</strong>g.<br />

Given <strong>the</strong> dynamism of rivers, it seems obvious that <strong>the</strong><br />

outcomes of restoration projects are uncerta<strong>in</strong>. However,<br />

<strong>the</strong> restoration community seems hesitant to admit that <strong>the</strong><br />

goals and science that restoration are founded upon are<br />

uncerta<strong>in</strong> too (Stewardson and Ru<strong>the</strong>rfurd, see Chapter 5).<br />

Aside from <strong>in</strong>direct references to uncerta<strong>in</strong>ty <strong>in</strong> adaptive<br />

management programs, <strong>the</strong> river management community<br />

has largely brushed uncerta<strong>in</strong>ties aside (Clark, 2002;<br />

Wissmar and Bisson, 2003c). It is unclear whe<strong>the</strong>r this is<br />

a conscious or passive decision, though <strong>in</strong>dividual decisions<br />

to ignore uncerta<strong>in</strong>ty can be plausibly attributed to<br />

one or more of <strong>the</strong> follow<strong>in</strong>g:<br />

ignorance of uncerta<strong>in</strong>ty and/or its signifi cance;<br />

<strong>the</strong> hope that uncerta<strong>in</strong>ty is <strong>in</strong>signifi cant;<br />

an acknowledgement of uncerta<strong>in</strong>ty, but not know<strong>in</strong>g<br />

how to deal with it;<br />

be<strong>in</strong>g mis<strong>in</strong>formed about uncerta<strong>in</strong>ty, lead<strong>in</strong>g to <strong>the</strong><br />

assumption that it is <strong>in</strong>signifi cant;<br />

be<strong>in</strong>g knowledgeable about uncerta<strong>in</strong>ty, but hav<strong>in</strong>g<br />

established its <strong>in</strong>signifi cance.


Newson and Clark (see Chapter 14) attribute <strong>the</strong> river<br />

manager’s current treatment of uncerta<strong>in</strong>ty to a ‘riskaverse’<br />

management culture that prefers to entrench itself<br />

<strong>in</strong> ‘rituals of verifi cation’ aimed at m<strong>in</strong>imis<strong>in</strong>g liability<br />

(Power, 1999). <strong>Uncerta<strong>in</strong>ty</strong> is also frequently misunderstood<br />

by <strong>the</strong> general public (Pollack, 2003; Riebeek, 2002)<br />

as someth<strong>in</strong>g negative and undesirable (Newson and<br />

Clark, see Chapter 14). A widespread misconception that<br />

science embodies certa<strong>in</strong> knowledge persists <strong>in</strong> <strong>the</strong> reports<br />

of <strong>the</strong> ma<strong>in</strong>stream media and views of <strong>the</strong> general public<br />

(Clark, 2002; Riebeek, 2002). Such misconceptions fuel<br />

expectations that science-based approaches to river restoration<br />

will yield positive outcomes. Ironically, people confront<br />

uncerta<strong>in</strong>ties everyday without hostility and choose<br />

to rout<strong>in</strong>ely make decisions about <strong>the</strong> future (Pollack,<br />

2003).<br />

<strong>Restoration</strong> science and <strong>the</strong> restoration literature are not<br />

much fur<strong>the</strong>r along than practitioners and decision makers.<br />

Wissmar and Bisson (2003b) asserted that ‘a better understand<strong>in</strong>g<br />

of variability and uncerta<strong>in</strong>ty is critical to <strong>the</strong><br />

successful implementation of restoration programs for<br />

aquatic and riparian systems.’ Yet, buried with<strong>in</strong> a rich<br />

literature on restoration are only occasional pass<strong>in</strong>g<br />

mentions of uncerta<strong>in</strong>ty (Brookes and Shields, 1996) and<br />

a handful of explicit treatments (Johnson and Brown,<br />

2001; Johnson et al., 2002; Johnson and R<strong>in</strong>aldi, 1997;<br />

Johnson and R<strong>in</strong>aldi, 1998; Wissmar and Bisson, 2003c).<br />

These studies understandably tend to focus on a specifi c<br />

type of uncerta<strong>in</strong>ty that might be reasonably articulated<br />

with<strong>in</strong> a specifi ed page limit, so a more holistic treatment<br />

of uncerta<strong>in</strong>ty is necessary (Newson and Clark, see<br />

Chapter 14; Van Asselt, 2000). <strong>Restoration</strong> is established<br />

as one important component of environmental management.<br />

It would be a shame to lose what public support<br />

already exists for restoration if political scrut<strong>in</strong>y recasts<br />

unrealistic expectations of river restoration as a ‘failure’,<br />

as opposed to <strong>the</strong> <strong>in</strong>adequate consideration of uncerta<strong>in</strong>ty<br />

<strong>the</strong>y truly stem from.<br />

3.2 WHAT DO WE MEAN BY UNCERTAINTY?<br />

3.2.1 A Lexicon of <strong>Uncerta<strong>in</strong>ty</strong><br />

In <strong>the</strong> simplest sense, uncerta<strong>in</strong>ty is a lack of sureness<br />

about someth<strong>in</strong>g or someone (Merriam-Webster, 1994).<br />

However, uncerta<strong>in</strong>ty can be more than simply a lack<br />

of knowledge. It persists even <strong>in</strong> areas where knowledge<br />

is extensive; and knowledge does not necessarily equate<br />

to truth or certa<strong>in</strong>ty (Van Asselt and Rotmans, 2002).<br />

There are at least 24 potential synonyms for <strong>the</strong> noun<br />

uncerta<strong>in</strong>ty and 27 synonyms for <strong>the</strong> adjective uncerta<strong>in</strong><br />

The Scope of Uncerta<strong>in</strong>ties <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 23<br />

(Table 3.1). There are a number of concepts related to and<br />

<strong>in</strong>fl uenced by uncerta<strong>in</strong>ty, but which differ from uncerta<strong>in</strong>ty<br />

itself. A selection of <strong>the</strong>se concepts is considered<br />

briefl y below.<br />

Accuracy: Accuracy refers to correctness or freedom from<br />

error. In measurement, accuracy refers to how close an<br />

<strong>in</strong>dividual measurement is to <strong>the</strong> ‘true’ or ‘correct’ value<br />

(Brown et al., 1994). The classic accuracy analogy is <strong>the</strong><br />

location of darts on a dart board – <strong>the</strong> closer <strong>the</strong> darts are<br />

to <strong>the</strong> <strong>in</strong>tended position (bull’s-eye) <strong>the</strong> more accurate. If<br />

one can be certa<strong>in</strong> about both <strong>the</strong> ‘true’ value (e.g. <strong>the</strong><br />

position of <strong>the</strong> bull’s-eye) and <strong>the</strong> value of <strong>the</strong> <strong>in</strong>dividual<br />

measurement (e.g. <strong>the</strong> position of <strong>the</strong> dart), <strong>the</strong>n <strong>the</strong> accuracy<br />

is actually a certa<strong>in</strong>ty. In practice, accuracy statements<br />

are uncerta<strong>in</strong> because ‘true’ values are often<br />

assumed and measurements have limited precision.<br />

Table 3.1 Potential synonyms of <strong>the</strong> noun ‘<strong>Uncerta<strong>in</strong>ty</strong>’ and<br />

<strong>the</strong> adjective ‘Uncerta<strong>in</strong>’<br />

Synonyms of <strong>Uncerta<strong>in</strong>ty</strong> Synonyms of Uncerta<strong>in</strong><br />

Ambiguity Ambiguous<br />

Indeterm<strong>in</strong>acy Causeless<br />

Capriciousness Capricious<br />

Chance Probabilistic<br />

– Deferred<br />

Danger Dangerous<br />

Disbelief Disbeliev<strong>in</strong>g<br />

Equivocation Equivocal<br />

Doubt Doubtful<br />

– Erratic<br />

Expectation –<br />

Future condition –<br />

Hesitation Hesitant<br />

Ignorance Ignorant<br />

Improbability Improbable<br />

Indecision Indecisive<br />

Indeterm<strong>in</strong>acy Indeterm<strong>in</strong>ant<br />

Insecurity Insecure<br />

Irresolution –<br />

Obscurity Obscure<br />

Surprise Surpris<strong>in</strong>g<br />

– Unau<strong>the</strong>ntic<br />

Un<strong>in</strong>telligibility Un<strong>in</strong>telligible<br />

– Unexpla<strong>in</strong>ed<br />

– Questionable<br />

Vacillation Vacillat<strong>in</strong>g<br />

Vagueness Vague<br />

– Undecided<br />

Unsureness Unsure<br />

Unpredictability Unpredictable


24 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Confi dence: Confi dence (e.g. <strong>in</strong> a statement, hypo<strong>the</strong>sis,<br />

measurement, feel<strong>in</strong>g or notion) relates to <strong>the</strong> degree of<br />

belief or level of certa<strong>in</strong>ty. Confi dence levels, for example,<br />

describe <strong>the</strong> probability that a given population parameter<br />

estimate falls with<strong>in</strong> a designated cont<strong>in</strong>uous statistical<br />

confi dence <strong>in</strong>terval.<br />

Divergence: Divergence describes a situation when similar<br />

causes produce dissimilar effects (Schumm, 1991). Divergence<br />

relates to uncerta<strong>in</strong>ty <strong>in</strong> situations where problems<br />

of cause and process are under consideration.<br />

Error: Error is <strong>the</strong> difference between a measured or calculated<br />

value and a ‘true’ value. In every day conversation,<br />

an error is a mistake. In science, error is <strong>the</strong> metric by<br />

which accuracy is reported and is not a synonym for uncerta<strong>in</strong>ty<br />

(Ellison et al., 2000). A ‘true’ value is certa<strong>in</strong> by<br />

defi nition. If <strong>the</strong> error between <strong>the</strong> ‘true’ value and a<br />

measured or calculated value is known <strong>the</strong>re is no uncerta<strong>in</strong>ty<br />

<strong>in</strong> pr<strong>in</strong>ciple. However, <strong>in</strong> practice ‘true’ values are<br />

often not known and <strong>in</strong>stead are assumed to be so, while<br />

<strong>the</strong> measured or calculated value may be uncerta<strong>in</strong>. Hence<br />

error becomes representative of uncerta<strong>in</strong>ty. Once errors<br />

are calculated, it can be helpful to consider whe<strong>the</strong>r <strong>the</strong><br />

error is systematic or random. Systematic errors stem from<br />

consistent mistakes and are often constant or predictable,<br />

affect<strong>in</strong>g <strong>the</strong> mean of a sample (i.e. bias, Trochim, 2000).<br />

Systematic errors can potentially be constra<strong>in</strong>ed as <strong>the</strong>ir<br />

source is identifi able. By contrast, random errors <strong>in</strong>fl uence<br />

<strong>the</strong> variability of a sample (not <strong>the</strong> mean) and are generally<br />

unpredictable or unconstra<strong>in</strong>able (Trochim, 2000).<br />

Exactness: Exactness is really a synonym for accuracy.<br />

However, it is worth po<strong>in</strong>t<strong>in</strong>g out that exactness has quite<br />

a different mean<strong>in</strong>g to exact. Exact statements or exact<br />

numbers, <strong>in</strong> pr<strong>in</strong>ciple, have no uncerta<strong>in</strong>ty about <strong>the</strong>m.<br />

They are statements of truth. By contrast, exactness is a<br />

relative measurement assigned to <strong>in</strong>exact statements or<br />

values (i.e. those with some uncerta<strong>in</strong>ty).<br />

Expectation: Expectation has to do with anticipation of<br />

probable or certa<strong>in</strong> events. <strong>Uncerta<strong>in</strong>ty</strong> fundamentally<br />

relates to expectations. When uncerta<strong>in</strong>ties are unknown,<br />

not fully considered or ignored, <strong>the</strong> degree that expectations<br />

may be unrealistic will generally <strong>in</strong>crease.<br />

Equifi nality: Equifi nality (also referred to as convergence),<br />

arises when different processes and causes produce<br />

similar effects (Schumm, 1991). In a modell<strong>in</strong>g context,<br />

Beven (1996a; 1996b) suggests that ‘<strong>the</strong> consequences of<br />

equifi nality are uncerta<strong>in</strong>ty <strong>in</strong> <strong>in</strong>ference and prediction.’ In<br />

a social context, a potentially limitless range of possibilities<br />

may lead to a s<strong>in</strong>gle event, such as <strong>the</strong> election or<br />

defeat of a politician.<br />

Precision: Precision is a measure of how closely <strong>in</strong>dividual<br />

measurements or calculations match one ano<strong>the</strong>r<br />

(Brown et al., 1994). Recall<strong>in</strong>g <strong>the</strong> dart board analogy<br />

from accuracy, a precisely thrown set of darts will cluster<br />

around one ano<strong>the</strong>r, but may be nowhere near <strong>the</strong> bull’seye.<br />

In measurement, <strong>the</strong> precision of an <strong>in</strong>strument refers<br />

to <strong>the</strong> fi nest scalar unit <strong>the</strong> <strong>in</strong>strument can resolve. Precision<br />

is related to uncerta<strong>in</strong>ty <strong>in</strong> that it defi nes a detection<br />

threshold, below which differences can not be discerned.<br />

Repeatability: Repeatability can be viewed as ei<strong>the</strong>r <strong>the</strong><br />

ability to reproduce <strong>the</strong> same measurement, result or<br />

calculation or <strong>the</strong> variability <strong>in</strong> repeated measurements,<br />

results or calculations. <strong>Uncerta<strong>in</strong>ty</strong> can simply limit<br />

repeatability or <strong>in</strong>crease variability.<br />

Risk: Risk is a measure of likelihood that an undesirable<br />

event or hazard will occur (Merriam-Webster, 1994). Ward<br />

(1998) credited Knight (1921) for mak<strong>in</strong>g <strong>the</strong> important<br />

clarifi cation between risk and <strong>the</strong> type of uncerta<strong>in</strong>ty for<br />

which <strong>the</strong>re exists ‘no valid basis of any k<strong>in</strong>d for classify<strong>in</strong>g<br />

<strong>in</strong>stances’:<br />

‘He used <strong>the</strong> term “risk” for situations <strong>in</strong> which an<br />

<strong>in</strong>dividual may not know <strong>the</strong> outcome of an event,<br />

but can form realistic expectations of <strong>the</strong> probabilities<br />

of <strong>the</strong> various possible outcomes based ei<strong>the</strong>r on<br />

ma<strong>the</strong>matical calculations or <strong>the</strong> history of previous<br />

occurrences.’<br />

Newson and Clark (see Chapter 14) contrast risk (with<br />

‘known’ impacts and probabilities) with uncerta<strong>in</strong>ty (with<br />

‘known’ impacts but ‘unknown’ probabilities) and ignorance<br />

(with ‘unknown’ impacts and probabilities).<br />

It is worth not<strong>in</strong>g that uncerta<strong>in</strong>ty itself and all <strong>the</strong><br />

related concepts outl<strong>in</strong>ed above are described <strong>in</strong> terms of<br />

<strong>the</strong>ir ‘degree’. That is, none of <strong>the</strong>se concepts are simple<br />

Aristotelian two-valued logic concepts (e.g. true–false).<br />

Each concept is measured along a cont<strong>in</strong>uum of values<br />

with end-members of total uncerta<strong>in</strong>ty (complete irreducible<br />

ignorance) and absolute certa<strong>in</strong>ty. Probabilistic uncerta<strong>in</strong>ty<br />

is an example of a quantifi cation of uncerta<strong>in</strong>ty, yet<br />

not all uncerta<strong>in</strong>ty is quantifi able. To quantify uncerta<strong>in</strong>ty<br />

it is necessary to estimate <strong>the</strong> degree of our limited knowledge.<br />

Yet, if a condition of irreducible ignorance is considered<br />

as one extreme of uncerta<strong>in</strong>ty, it is diffi cult at best<br />

to estimate <strong>the</strong> degree of someth<strong>in</strong>g we do not even know<br />

exists. With<strong>in</strong> this broad view of uncerta<strong>in</strong>ty, uncerta<strong>in</strong>ty<br />

might also be considered along a cont<strong>in</strong>uum that refl ects<br />

our ability to quantify it (Figure 3.1).<br />

In summary, when someone mentions uncerta<strong>in</strong>ty casually,<br />

it is diffi cult to discern whe<strong>the</strong>r <strong>the</strong>y are referr<strong>in</strong>g to


limited knowledge, a lack of knowledge altoge<strong>the</strong>r, or one<br />

of <strong>the</strong> above-mentioned concepts that are <strong>in</strong>fl uenced by<br />

uncerta<strong>in</strong>ty. Moreover, <strong>the</strong> lexicon provided here conta<strong>in</strong>s<br />

concepts that are highly <strong>in</strong>ter-related and easily confused.<br />

Similar to vague, pseudo-scientifi c buzzwords and catchall<br />

phrases like holistic and <strong>in</strong>tegrated, <strong>the</strong> term ‘uncerta<strong>in</strong>ty’<br />

alone evidently has little mean<strong>in</strong>g until its details<br />

are unravelled.<br />

3.2.2 A Typology for <strong>Uncerta<strong>in</strong>ty</strong><br />

S<strong>in</strong>ce uncerta<strong>in</strong>ty is so hard to defi ne, a classifi cation of<br />

uncerta<strong>in</strong>ty is often used (Van Asselt and Rotmans, 2002).<br />

The utility of any typology or classifi cation is ultimately<br />

dependent on its application (Kondolf, 1995b; Lew<strong>in</strong>,<br />

2001). Rotmans and van Asselt (2001) astutely po<strong>in</strong>ted<br />

out ‘<strong>the</strong>re is not one overall typology that satisfactorily<br />

covers all sorts of uncerta<strong>in</strong>ties, but that <strong>the</strong>re are many<br />

possible typologies’. In <strong>the</strong> context of this review, a<br />

typology was sought which considered sources of uncerta<strong>in</strong>ty<br />

and did not unnecessarily ignore any type of<br />

uncerta<strong>in</strong>ty. The exist<strong>in</strong>g van Asselt (2000) typology was<br />

chosen over o<strong>the</strong>rs because of its generic and <strong>in</strong>clusive<br />

consideration of uncerta<strong>in</strong>ty. The typology was fi rst <strong>in</strong>troduced<br />

<strong>in</strong> detail <strong>in</strong> van Asselt (2000) and concisely reviewed<br />

<strong>in</strong> Rotmans and van Asselt (2001) and van Asselt and<br />

Rotmans (2002).<br />

At <strong>the</strong> highest level, two sources of uncerta<strong>in</strong>ty exist:<br />

uncerta<strong>in</strong>ty due to variability and uncerta<strong>in</strong>ty due to<br />

limited knowledge (Figure 3.2). Van Asselt and Rotmans<br />

(2002) presented uncerta<strong>in</strong>ty due to variability fi rst as<br />

<strong>the</strong>se uncerta<strong>in</strong>ties ultimately comb<strong>in</strong>e to contribute to<br />

uncerta<strong>in</strong>ty due to limited knowledge. Environmental<br />

management is concerned with <strong>the</strong> management of <strong>in</strong>herently<br />

variable natural and managed systems. Knowledge<br />

The Scope of Uncerta<strong>in</strong>ties <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 25<br />

Figure 3.1 The quantifi able cont<strong>in</strong>uum of uncerta<strong>in</strong>ty (Once uncerta<strong>in</strong>ties are acknowledge as unquantifi ed uncerta<strong>in</strong>ties, <strong>in</strong>creased<br />

knowledge about <strong>the</strong> uncerta<strong>in</strong>ties will determ<strong>in</strong>e <strong>the</strong>ir position on <strong>the</strong> cont<strong>in</strong>uum.)<br />

about natural change and variability <strong>in</strong> ecosystems, fl uvial<br />

systems and hydrologic systems is <strong>in</strong>complete and hence<br />

contributes to uncerta<strong>in</strong>ty due to limited knowledge<br />

(Wissmar and Bisson, 2003a). Five dist<strong>in</strong>ct subclasses<br />

of uncerta<strong>in</strong>ty due to variability are proposed: <strong>in</strong>herent<br />

natural randomness, value diversity (socio-political),<br />

behavioural diversity, societal randomness and technological<br />

surprise. Inherent natural randomness is attributed to<br />

‘<strong>the</strong> nonl<strong>in</strong>ear, chaotic and unpredictable nature of natural<br />

processes’. The natural variability of river systems should<br />

be a fundamental consideration <strong>in</strong> <strong>in</strong>tegrated river bas<strong>in</strong><br />

management and restoration; it is reviewed thoroughly <strong>in</strong><br />

Wissmar and Bisson (2003c). Value diversity, behavioural<br />

diversity and societal randomness each contribute to<br />

uncerta<strong>in</strong>ties <strong>in</strong> environmental management, particularly<br />

through stakeholder negotiations, public support, project<br />

fund<strong>in</strong>g, policy mak<strong>in</strong>g and <strong>in</strong>dividual perspectives. Technological<br />

surprises result from new breakthroughs, which<br />

may provide unforeseen benefi ts and/or br<strong>in</strong>g unforeseen<br />

consequences.<br />

Van Asselt and Rotmans (2002) separated seven types<br />

of uncerta<strong>in</strong>ty due to limited knowledge. Unlike uncerta<strong>in</strong>ties<br />

due to variability, <strong>the</strong>se are thought to map out<br />

along a cont<strong>in</strong>uum that refl ects <strong>the</strong> relative degree of<br />

uncerta<strong>in</strong>ty. At <strong>the</strong> highest degree of uncerta<strong>in</strong>ty are four<br />

‘structural uncerta<strong>in</strong>ties’ (van Asselt and Rotmans,<br />

2002):<br />

Irreducible ignorance: ‘We cannot know.’<br />

Indeterm<strong>in</strong>acy: ‘We will never know.’<br />

Reducible ignorance: ‘We do not know what we do not<br />

know.’<br />

Confl ict<strong>in</strong>g evidence: Knowledge is not fact but <strong>in</strong>terpretation,<br />

and <strong>in</strong>terpretations frequently contradict and<br />

challenge each o<strong>the</strong>r. ‘We don’t know what we know.’


26 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Figure 3.2 Typology for sources and degree of uncerta<strong>in</strong>ty (Adapted from Van Asselt’s (2000) proposed typology for uncerta<strong>in</strong>ties<br />

<strong>in</strong> <strong>in</strong>tegrated assessment.)<br />

Van Asselt and Rotmans (2002) <strong>the</strong>n proposed a transition<br />

<strong>in</strong>to ‘unreliability’ uncerta<strong>in</strong>ties of a relatively lesser<br />

degree:<br />

Practically immeasurable: A lack of data or <strong>in</strong>formation<br />

is always a reality <strong>in</strong> study<strong>in</strong>g natural systems.<br />

Not only are many natural phenomena <strong>in</strong>credibly diffi<br />

cult or impossible to measure, all are fundamentally<br />

limited by problems of temporal and spatial resolution,<br />

up-scal<strong>in</strong>g and averag<strong>in</strong>g (Kavvas, 1999). ‘We<br />

know what we don’t know’ (Van Asselt and Rotmans,<br />

2002).<br />

Lack of Observations and Measurements: Although<br />

<strong>in</strong> pr<strong>in</strong>ciple this is easy to identify and augment, <strong>in</strong><br />

practice this is always a factor. Borrow<strong>in</strong>g from van<br />

Asselt and Rotmans (2002): ‘could have, should<br />

have, would have, but didn’t.’<br />

Inexactness: Related to lack of precision, lack of<br />

accuracy, measurement and calculation errors. Under<br />

Klir and Yuan’s (1995) typology, <strong>the</strong>se are considered<br />

‘fuzz<strong>in</strong>ess’ or vagueness.<br />

The van Asselt (2000) typology is both more general<br />

and detailed than o<strong>the</strong>r typologies such as Klir and Yuan<br />

(1995). However, all provide a reasonable means to deal<br />

with <strong>the</strong> fi rst step to understand<strong>in</strong>g uncerta<strong>in</strong>ty. Namely,<br />

<strong>the</strong>y allow a systematic identifi cation of sources and types<br />

of uncerta<strong>in</strong>ties that could work <strong>in</strong> ei<strong>the</strong>r <strong>in</strong>dividual river<br />

restoration projects or <strong>in</strong>ternational policy mak<strong>in</strong>g on<br />

water and environmental management (see also Chapters<br />

1 and 2). In practice, it is recognised that <strong>the</strong> semantics of<br />

uncerta<strong>in</strong>ty will always be <strong>in</strong>terpreted differently <strong>in</strong> different<br />

professional contexts (Newson and Clark, see Chapter<br />

14). However, with<strong>in</strong> <strong>the</strong> context of this chapter, <strong>the</strong> van<br />

Asselt (2000) typology and associated mean<strong>in</strong>gs will be<br />

used consistently.<br />

3.2.3 How do Knowledge and <strong>Uncerta<strong>in</strong>ty</strong> Relate?<br />

The positivist view (Van Asselt and Rotmans, 2002) contends<br />

that as knowledge <strong>in</strong>creases, uncerta<strong>in</strong>ty decreases.<br />

Brookes et al. (1998) made <strong>the</strong> more restrictive but<br />

contradictory generalisation that ‘as knowledge relat<strong>in</strong>g<br />

to rivers and <strong>the</strong>ir fl oodpla<strong>in</strong>s <strong>in</strong>creases, uncerta<strong>in</strong>ty is<br />

<strong>in</strong>creased ra<strong>the</strong>r than decreased.’ So, which is it? In reality,<br />

<strong>the</strong>re is no unique relationship between uncerta<strong>in</strong>ty and<br />

knowledge (Van Asselt and Rotmans, 2002), nor is uncerta<strong>in</strong>ty<br />

a fi xed quantity that will always be reduced by scientifi<br />

c research (Jamieson, 1996). It is a highly contextual<br />

relationship dependent on <strong>the</strong> type of uncerta<strong>in</strong>ty (i.e.<br />

uncerta<strong>in</strong>ty due to lack of knowledge versus variability)<br />

and <strong>the</strong> specifi c circumstances under consideration. A few


examples of potential relationships between knowledge<br />

and uncerta<strong>in</strong>ty us<strong>in</strong>g <strong>the</strong> nomenclature of <strong>the</strong> van<br />

Asslet typology are illustrated <strong>in</strong> Figure 3.3. Hav<strong>in</strong>g<br />

established <strong>the</strong> basic term<strong>in</strong>ology of uncerta<strong>in</strong>ty, it is<br />

possible to discuss <strong>the</strong> sources of uncerta<strong>in</strong>ty with<strong>in</strong><br />

river restoration.<br />

The Scope of Uncerta<strong>in</strong>ties <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 27<br />

3.3 REVISITING RIVER RESTORATION<br />

AND UNCERTAINTY<br />

It is diffi cult to generalise about <strong>the</strong> importance of uncerta<strong>in</strong>ty<br />

simply because restoration activities and <strong>the</strong> restoration<br />

community itself are so diverse. The stakeholders who<br />

Figure 3.3 Some potential relationships between knowledge and uncerta<strong>in</strong>ty through time (Contrary to <strong>the</strong> argument of <strong>the</strong> positivist,<br />

no unique <strong>in</strong>verse relationship between uncerta<strong>in</strong>ty and knowledge exists.)


28 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

<strong>in</strong>itiate river restoration projects <strong>in</strong>clude private <strong>in</strong>dividuals,<br />

non-governmental organisations (NGOs), governmental<br />

organisations and various collaborative comb<strong>in</strong>ations<br />

of <strong>the</strong> above. The restoration community is also comprised<br />

of practitioners, decision makers and scientists. No attempt<br />

is made here to list ‘all’ <strong>the</strong> uncerta<strong>in</strong>ties encountered<br />

throughout <strong>the</strong> restoration process as <strong>the</strong> daunt<strong>in</strong>g list<br />

would never be comprehensive, and is entirely perspective<br />

and project specifi c. For example, <strong>the</strong>re is little consensus<br />

over <strong>the</strong> mean<strong>in</strong>g of <strong>the</strong> term ‘river restoration’ with at<br />

least 30 different authors propos<strong>in</strong>g different defi nitions<br />

(Lemons and Victor, see Chapter 1; NAP, 2002; Newson,<br />

2002; Sear, 1994; Stockwell, 2000). Similar to Shields<br />

et al. (2003), ‘river restoration’ <strong>in</strong> this book is used as a<br />

catch-all term for a variety of management responses and<br />

activities used to address perceived problems with rivers<br />

(Kondolf, 1996). As a start<strong>in</strong>g po<strong>in</strong>t, a generic decision<br />

process, which most restoration projects loosely follow,<br />

highlight<strong>in</strong>g some of <strong>the</strong> common sources of uncerta<strong>in</strong>ty<br />

is mapped out <strong>in</strong> Table 3.2.<br />

3.3.1 Motives for <strong>Restoration</strong><br />

Once river restoration projects ga<strong>in</strong> momentum, it is easy<br />

to lose sight of why <strong>the</strong>y were orig<strong>in</strong>ally envisioned<br />

(Stewardson and Ru<strong>the</strong>rfurd, see Chapter 5). Here, <strong>the</strong><br />

motives for restoration are considered to represent more<br />

generalised aims than formalised and specifi c restoration<br />

objectives and activities (i.e. <strong>the</strong> ‘why’ <strong>in</strong>stead of <strong>the</strong><br />

‘what’). Eight common types of motives for river restoration<br />

(still o<strong>the</strong>rs exist) are listed below:<br />

1. Ecosystem <strong>Restoration</strong><br />

2. Habitat <strong>Restoration</strong><br />

3. Flood Control/Defence<br />

4. Floodpla<strong>in</strong> Reconnection<br />

5. Property and Infrastructure Protection (bank<br />

stability)<br />

6. Sediment Management<br />

7. Water Quality<br />

8. Aes<strong>the</strong>tic and Recreational.<br />

Considerable overlap exists between many of <strong>the</strong> above.<br />

For example, fl oodpla<strong>in</strong> reconnection can be a type of<br />

fl ood control. Habitat restoration and water quality restoration<br />

are sometimes considered forms of ecosystem<br />

restoration. In ano<strong>the</strong>r example, water quality restoration<br />

could be viewed by some as sediment management or<br />

by o<strong>the</strong>rs as aes<strong>the</strong>tic or recreational restoration. Thus, a<br />

hierarchical organisation of restoration motives would be<br />

highly subjective and dependent on <strong>in</strong>dividual values and<br />

perspectives. This <strong>in</strong> itself is not necessarily problematic.<br />

However, it represents a form of communication uncerta<strong>in</strong>ty<br />

aris<strong>in</strong>g out of value diversity which is often taken<br />

for granted. Once <strong>the</strong> motives (why to do it) for restoration<br />

are established, restoration aims fall <strong>in</strong>to place, but more<br />

specifi c objectives (what and how to do it) require careful<br />

consideration.<br />

Table 3.2 Sources of uncerta<strong>in</strong>ty <strong>in</strong> an environmental management decision process structure (Adapted from Chapman & Ward<br />

(2002))<br />

Stage <strong>in</strong> Decision Process <strong>Uncerta<strong>in</strong>ty</strong> About<br />

Monitor <strong>the</strong> environment and current operations Completeness, veracity and accuracy of <strong>in</strong>formation received, mean<strong>in</strong>g<br />

with<strong>in</strong> <strong>the</strong> organisation<br />

of <strong>in</strong>formation, <strong>in</strong>terpretation of implications<br />

Recognise an issue Signifi cance of issue, urgency, need for action<br />

Scope <strong>the</strong> Decision Appropriate frame of reference, scope of relevant organisation activities,<br />

who is <strong>in</strong>volved, who should be <strong>in</strong>volved, extent of separation from<br />

o<strong>the</strong>r decision issues<br />

Determ<strong>in</strong>e <strong>the</strong> performance criteria Relevant performance criteria, whose criteria, appropriate metrics,<br />

appropriate priorities and trade offs between different criteria<br />

Identify alternative courses of action† Nature of alternatives available (scope, tim<strong>in</strong>g, logistics <strong>in</strong>volved), what<br />

is possible, level of detail required, time available to identify<br />

alternatives<br />

Predict <strong>the</strong> outcomes of courses of action† Consequences, nature of <strong>in</strong>fl uenc<strong>in</strong>g factors, size of <strong>in</strong>fl uenc<strong>in</strong>g factors,<br />

effects and <strong>in</strong>teractions between <strong>in</strong>fl uenc<strong>in</strong>g factors (variability and<br />

tim<strong>in</strong>g), nature and signifi cance of assumptions made<br />

Choose a course of action How to weigh and compare predicted outcomes<br />

Implement <strong>the</strong> chosenalternative* How alternatives will work <strong>in</strong> practice<br />

Monitor and reviewperformance‡ What to monitor, how often to monitor, when to take fur<strong>the</strong>r action<br />

† = Most decision support systems only provide <strong>in</strong>put at <strong>the</strong>se levels; * = The precautionary pr<strong>in</strong>ciple is implemented here; ‡ = Adaptive management<br />

starts here and feeds back through <strong>the</strong> process as necessary.


Many have argued that uncerta<strong>in</strong>ty <strong>in</strong> assess<strong>in</strong>g restoration<br />

success arises from <strong>in</strong>adequate, vague and unclear<br />

restoration objectives (Jungwirth et al., 2002; Kondolf,<br />

1995a; Sk<strong>in</strong>ner et al., see Chapter 10). Motives may serve<br />

well as aims (not necessarily to be achieved by an <strong>in</strong>dividual<br />

project) but <strong>the</strong>y are <strong>in</strong>suffi cient to act as detailed<br />

project objectives, which <strong>in</strong> pr<strong>in</strong>ciple should be achievable.<br />

Us<strong>in</strong>g restoration motives carelessly as objectives<br />

produces unrealistic expectations. For example, <strong>in</strong> a recent<br />

request for proposals to fund community-based river restoration<br />

projects by American <strong>River</strong>s and <strong>the</strong> National<br />

Oceanic and Atmospheric Association, applicants were<br />

asked to demonstrate that <strong>the</strong>ir project: will successfully<br />

restore anadromous fi sh habitat, access to exist<strong>in</strong>g<br />

anadromous fi sh habitat, or natural river<strong>in</strong>e functions; is<br />

<strong>the</strong> correct approach, based on ecological, social, economic,<br />

and eng<strong>in</strong>eer<strong>in</strong>g considerations; will m<strong>in</strong>imise any<br />

identifi able short or long term negative impacts to <strong>the</strong> river<br />

system as a result of <strong>the</strong> project . . .’<br />

The problem with requir<strong>in</strong>g an applicant to make such<br />

bold statements about <strong>in</strong>dividual projects is that it asserts<br />

a level of confi dence <strong>in</strong> restoration simply not warranted<br />

by current science or practice and creates unrealistic<br />

expectations 4 (Stewardson and Ru<strong>the</strong>rfurd, see Chapter 5).<br />

Subtly reword<strong>in</strong>g such requirements to account for uncerta<strong>in</strong>ty<br />

could help recast river restoration <strong>in</strong> a tone commensurate<br />

with our abilities and uncerta<strong>in</strong>ties. Interest<strong>in</strong>gly,<br />

<strong>the</strong>se objectives are consistent with Clark’s (2002) critical<br />

synopsis of Predictive Management as opposed to<br />

adaptive management as <strong>the</strong> current model <strong>in</strong> river<br />

management.<br />

The restoration community has burdened itself with <strong>the</strong><br />

idea that restoration objectives should be scientifi cally<br />

based (Davis and Slobodk<strong>in</strong>, 2004). While science surely<br />

has an important role <strong>in</strong> restoration, Davis and Slobodk<strong>in</strong><br />

(2004) argued that determ<strong>in</strong><strong>in</strong>g restoration objectives<br />

is fundamentally a value-based and subjective process.<br />

Noth<strong>in</strong>g is seen as <strong>in</strong>herently wrong with this reality, so<br />

long as it is transparently recognised. From an uncerta<strong>in</strong>ty<br />

perspective, this means that restoration objectives are<br />

<strong>the</strong>refore sources of uncerta<strong>in</strong>ty due to variability; namely<br />

value diversity, behavioural diversity and societal randomness.<br />

For example, <strong>the</strong> fate of 81 000 hectares of forest<br />

land allocated for ecosystem restoration around <strong>the</strong> city of<br />

Chicago, Ill<strong>in</strong>ois has pitted two ‘environmental’ groups<br />

aga<strong>in</strong>st each o<strong>the</strong>r based on <strong>the</strong>ir contrast<strong>in</strong>g notions of<br />

‘what is natural’. The divergent environmental views are<br />

essentially split between preservationists, who wish to preserve<br />

<strong>the</strong> forest land planted <strong>in</strong> <strong>the</strong> 1800s, and restoration-<br />

4 This is fundamentally a communication uncerta<strong>in</strong>ty result<strong>in</strong>g<br />

from socio-political value diversity (see Figure 3.1).<br />

The Scope of Uncerta<strong>in</strong>ties <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 29<br />

ists, who want to restore <strong>the</strong> pre-settlement (1830s) prairie<br />

and savannah (Alario and Brün, 2001). Both evoke emotional<br />

arguments, which can be supported on scientifi c<br />

grounds. ‘Which is right?’ is <strong>the</strong> wrong question to ask of<br />

science. Alario and Brün (2001) concluded that <strong>the</strong> appropriate<br />

arena to decide such an issue is a political decision<br />

mak<strong>in</strong>g process.<br />

3.3.2 Notions that Drive <strong>Restoration</strong><br />

Underly<strong>in</strong>g motives for river restoration and <strong>the</strong> eventual<br />

specifi c techniques tried to achieve <strong>the</strong>m are some very<br />

basic, yet highly uncerta<strong>in</strong> notions. S<strong>in</strong>ce <strong>the</strong>se basic<br />

notions are rarely questioned, it is important to highlight<br />

how <strong>the</strong>y <strong>in</strong>troduce uncerta<strong>in</strong>ty. Notions are also known<br />

as ‘Lietbilds’ – or target visions – and have ga<strong>in</strong>ed widespread<br />

acceptance <strong>in</strong> <strong>the</strong> restoration literature (Hughes,<br />

1995; Jungwirth et al., 2002; Kern, 1992). Notions, such<br />

as those <strong>in</strong> Table 3.3, that drive restoration strategies<br />

are frequently based on societal values and beliefs, or on<br />

popular, but by no means certa<strong>in</strong>, scientifi c paradigms<br />

(Davis and Slobodk<strong>in</strong>, 2004; McDonald et al., 2004;<br />

Rhoads et al., 1999).<br />

Falkenmark and Folke (2002) argued that susta<strong>in</strong>able<br />

catchment management must be based on ethical pr<strong>in</strong>ciples.<br />

They suggest that management based on scientifi c<br />

pr<strong>in</strong>ciples alone is primarily concerned with ‘do<strong>in</strong>g <strong>the</strong><br />

th<strong>in</strong>g right’, whereas notions that drive restoration strategies<br />

are actually driven by ‘do<strong>in</strong>g <strong>the</strong> right th<strong>in</strong>g.’ It is a<br />

presumption that good ethical practice generally translates<br />

<strong>in</strong>to good biological practice (Pister, 2001). Hence notions<br />

are vague ideas, perhaps based on scientifi c knowledge,<br />

but primarily supported by ethical beliefs and societal<br />

values. The restoration literature is rarely explicit <strong>in</strong> dist<strong>in</strong>guish<strong>in</strong>g<br />

<strong>the</strong> notions it advocates from <strong>the</strong> science used<br />

to support it. Phillip Williams (personal communication)<br />

asserts that ‘rigour’ <strong>in</strong> restoration plann<strong>in</strong>g should start<br />

with <strong>the</strong> development of an explicit conceptual model<br />

transparently describ<strong>in</strong>g our notions of how <strong>the</strong> river<br />

system functions5 . Such a conceptual model should identify<br />

both <strong>the</strong> historical context and <strong>the</strong> present day limitations<br />

(i.e. uncerta<strong>in</strong>ties). Wheaton et al. (2004a) argued<br />

that numerous conceptual models <strong>in</strong> <strong>the</strong> scientifi c literature<br />

already exist and can be borrowed or modifi ed to<br />

formulate a site or bas<strong>in</strong> specifi c conceptual model as <strong>the</strong><br />

basis for restoration. Yet, Stewardson and Ru<strong>the</strong>rfurd (see<br />

Chapter 5) describe three levels <strong>in</strong> restoration from which<br />

5 In pr<strong>in</strong>ciple, <strong>the</strong> process of ‘rigour’ <strong>in</strong> restoration plann<strong>in</strong>g still<br />

follows <strong>the</strong> generic environmental management decision process<br />

of Table 3.2. In essence what Phillip Williams, a seasoned practitioner,<br />

describes is an <strong>in</strong>formal Decision Support System<br />

(DSS).


30 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Table 3.3 Common motives that guide notions and drive river restoration efforts<br />

Notion Example(s)<br />

What is Natural?<br />

Nature is <strong>in</strong> Equilibrium ‘<strong>the</strong> equilibrium between sediment supply and available transport capacity.’ (Soar &<br />

Thorne 2001); ‘landforms can be considered as ei<strong>the</strong>r a stage <strong>in</strong> a cycle of erosion or as<br />

a system <strong>in</strong> dynamic equilibrium.’ (Schumm & Lichty 1965).<br />

Nature is <strong>in</strong> fl ux ‘Restored ecosystems are those <strong>in</strong> which <strong>the</strong> rates and types of disturbance do not exceed<br />

<strong>the</strong> capacity of <strong>the</strong> system to respond to <strong>the</strong>m.’ (Hruby 2003).<br />

Nature Constant ‘confi dence on global stability; <strong>the</strong>re are no limitations to development’ (Levy et al. 2000).<br />

Nature Balanced ‘<strong>the</strong> environment is forgiv<strong>in</strong>g of most shocks, but large perturbations can knock ecological<br />

variables <strong>in</strong>to new regions of <strong>the</strong> landscape.’ (Levy et al. 2000).<br />

Nature Ephemeral ‘<strong>the</strong> environment can not safely tolerate human modifi cations’ (Levy et al. 2000).<br />

Nature Resilient ‘ecosystems are adaptive, evolutionary, and self organis<strong>in</strong>g . . . ecological systems often<br />

thrive under conditions of high variability’ (Levy et al. 2000).<br />

Physical Integrity<br />

Physical Integrity ‘Physical Integrity for rivers refers to a set of active fl uvial processes and landforms<br />

where<strong>in</strong> channel, fl oodpla<strong>in</strong>s, sediments, and overall spatial confi guration ma<strong>in</strong>ta<strong>in</strong> a<br />

dynamic equilibrium, with adjustments not exceed<strong>in</strong>g limits of change defi ned by<br />

societal values. <strong>River</strong>s possess physical <strong>in</strong>tegrity when <strong>the</strong>ir processes and forms<br />

ma<strong>in</strong>ta<strong>in</strong> active connections with each o<strong>the</strong>r <strong>in</strong> <strong>the</strong> present hydrologic regime.’<br />

(Graf 2001).<br />

Alluvial <strong>River</strong> Attributes Several commonly known concepts that govern how alluvial channels work have been<br />

compiled <strong>in</strong>to a set of ‘attributes’ for alluvial river <strong>in</strong>tegrity (Trush et al. 2000).<br />

Ecological Integrity Ecological Integrity ‘ma<strong>in</strong>tenance of all <strong>in</strong>ternal and external processes and attributes<br />

<strong>in</strong>teract<strong>in</strong>g with <strong>the</strong> environment <strong>in</strong> such a way that <strong>the</strong> biotic community corresponds<br />

to <strong>the</strong> natural state of <strong>the</strong> type-specifi c aquatic habitat, accord<strong>in</strong>g to <strong>the</strong> pr<strong>in</strong>ciples of<br />

self-regulation, resilience and resistance.’ (Angermeier & Karr 1994).<br />

High Biodiversity = Ecological Natural systems foster biodiversity and artifi cial systems are homogenized and dom<strong>in</strong>ated<br />

Integrity<br />

by <strong>in</strong>vasive species (Ward et al. 2002, Lister 1998).<br />

Morphological Diversity =<br />

Biological Diversity<br />

epistemological uncerta<strong>in</strong>ties emerge: <strong>the</strong> validity of<br />

<strong>the</strong> conceptual model; whe<strong>the</strong>r <strong>the</strong> proposed <strong>in</strong>tervention<br />

results <strong>in</strong> <strong>the</strong> planned geomorphic change; and whe<strong>the</strong>r<br />

<strong>the</strong> change is susta<strong>in</strong>able. They <strong>the</strong>n caution that <strong>the</strong><br />

validity of <strong>the</strong> conceptual model is <strong>the</strong> source of <strong>the</strong> ‘most<br />

uncerta<strong>in</strong>ty.’ Return<strong>in</strong>g to Phillip William’s concept of<br />

rigour <strong>in</strong> plann<strong>in</strong>g, he argues restoration objectives should<br />

be based on an understand<strong>in</strong>g of how <strong>the</strong> conceptual<br />

model <strong>in</strong>teracts and responds to various societal motives<br />

Newson (2002) did not dispute <strong>the</strong> abundance of evidence support<strong>in</strong>g <strong>the</strong> l<strong>in</strong>kages between<br />

channel dynamics and biodiversity, but criticises <strong>the</strong> lack of direct collaboration<br />

between geomorphologists and ecologists to substantiate <strong>the</strong> l<strong>in</strong>ks <strong>in</strong> river management:<br />

‘<strong>the</strong> mantra “morphological diversity = biodiversity” currently rema<strong>in</strong>s an act of faith.’<br />

What is Susta<strong>in</strong>able?<br />

Susta<strong>in</strong>ability Accord<strong>in</strong>g to Cairns (2003), <strong>the</strong> notion of susta<strong>in</strong>ability is based on ‘<strong>the</strong> assumption that<br />

humank<strong>in</strong>d has <strong>the</strong> right to alter <strong>the</strong> planet so that human life can <strong>in</strong>habit Earth<br />

<strong>in</strong>defi nitely.’<br />

Geomorphic Susta<strong>in</strong>ability ‘susta<strong>in</strong>ability encompasses <strong>the</strong> notion of self-regulation of spontaneous functions (e.g.<br />

sediment deposition, colonisation and succession of vegetation) with m<strong>in</strong>imal<br />

<strong>in</strong>tervention and no adverse impact on <strong>the</strong> future aquatic environment whilst<br />

ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g <strong>the</strong> functions of <strong>the</strong> channel demanded by society (fl ood control,<br />

navigation etc.).’ (Sear 1996).<br />

(NRC, 1992). Based on specifi c objectives, a measurable<br />

set of <strong>in</strong>dicators and target levels can be selected (Doyle<br />

et al., 2000; Levy et al., 2000; Merkle and Kaupenjohann,<br />

2000; Smeets and Weter<strong>in</strong>gs, 1999). F<strong>in</strong>ally, a comparison<br />

of predicted <strong>in</strong>dicator responses to restoration <strong>in</strong>tervention<br />

versus <strong>in</strong>action should be used to decide whe<strong>the</strong>r restoration<br />

is appropriate. Although available science may be<br />

used to <strong>in</strong>form <strong>the</strong> steps lead<strong>in</strong>g up to this decision<br />

(Lemons and Victor, see Chapter 1), <strong>the</strong> decision whe<strong>the</strong>r


or not to proceed is ultimately a political one (Alario and<br />

Brun, 2001).<br />

3.3.3 Approaches to <strong>Restoration</strong><br />

Generally, river restoration projects consist of three<br />

components: plann<strong>in</strong>g, implementation and evaluation.<br />

The diversity of approaches available to implement <strong>the</strong>se<br />

components ra<strong>the</strong>r appropriately refl ects <strong>the</strong> varied types<br />

(motives) of restoration projects and physiographic sett<strong>in</strong>gs<br />

<strong>the</strong>y are applied <strong>in</strong>. Thus, historical and spatial cont<strong>in</strong>gencies<br />

are contribut<strong>in</strong>g to uncerta<strong>in</strong>ties due to natural<br />

variability (Phillips, 2001). Indeed, a plethora of restoration<br />

approaches and strategies has been formalised <strong>in</strong> both<br />

<strong>the</strong> peer-reviewed and grey literature (Wheaton et al.,<br />

2004a). Examples range from generalised approaches for<br />

stream restoration (e.g. FISRWG, 1998; Jungwirth et al.,<br />

2002; Koehn et al., 2001; NRC, 1992; RRC, 2002) to<br />

more specifi c strategies <strong>in</strong>corporat<strong>in</strong>g: fl uvial geomorphology<br />

(e.g. Brookes and Sear, 1996; Gilvear, 1999;<br />

Kondolf, 2000; Sear, 1994), ecosystem <strong>the</strong>ory (e.g.<br />

Richards et al., 2002; Stanford et al., 1996), hydraulic<br />

eng<strong>in</strong>eer<strong>in</strong>g (e.g. Shields, 1996) and detailed design procedures<br />

(Miller et al., 2001; Shields et al., 2003; Wheaton<br />

et al., 2004b). Most of <strong>the</strong> approaches have parallels <strong>in</strong><br />

structure and ideology (Wheaton et al., 2004a).<br />

Popular labels used to describe restoration approaches<br />

<strong>in</strong>clude holistic, science-based, <strong>in</strong>tegrated and multidiscipl<strong>in</strong>ary<br />

(Hildén, 2000; Jungwirth et al., 2002;<br />

Wissmar and Bisson, 2003a). S<strong>in</strong>ce most approaches<br />

purport or aim to be all of <strong>the</strong>se (Wheaton et al., 2004a),<br />

and <strong>the</strong> converse of each is perceived as negative, <strong>the</strong>re is<br />

little value <strong>in</strong> discrim<strong>in</strong>at<strong>in</strong>g approaches on <strong>the</strong>se grounds.<br />

However, <strong>the</strong>ir components (i.e. plann<strong>in</strong>g, implementation<br />

and monitor<strong>in</strong>g) can be differentiated us<strong>in</strong>g three descriptive<br />

metrics: <strong>the</strong> scale of restoration; form based versus<br />

process based; and active versus passive. These metrics<br />

can provide <strong>in</strong>sight <strong>in</strong>to <strong>the</strong> types of uncerta<strong>in</strong>ties encountered<br />

and expectations placed on restoration projects<br />

dur<strong>in</strong>g plann<strong>in</strong>g, implementation and monitor<strong>in</strong>g.<br />

S<strong>in</strong>ce <strong>the</strong> late 1990s, approaches almost unanimously<br />

call for catchment scale plann<strong>in</strong>g <strong>in</strong> restoration6 . However,<br />

confusion arises over whe<strong>the</strong>r this means: restore <strong>the</strong><br />

entire catchment; use watershed assessments to nest reach<br />

scale restoration <strong>in</strong> a catchment context (e.g. Bohn and<br />

Kershner, 2002; Brookes and Shields, 1996; Walker et al.,<br />

2002) or undertake a range of management and restoration<br />

activities across various spatial scales but nested with<strong>in</strong> a<br />

catchment context (e.g. Frissell et al., 1993; Roni et al.,<br />

2002). Ecosystem degradation has often taken place over<br />

6 See also Table 3.1.<br />

The Scope of Uncerta<strong>in</strong>ties <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 31<br />

many decades or centuries and extends across landscape,<br />

catchment and regional scales (Palmer et al., 1997).<br />

However, restor<strong>in</strong>g an entire catchment is rarely viable<br />

(Brookes and Shields, 1996). Even those who call for<br />

ecological restoration of <strong>the</strong> entire catchment (e.g. Frissell<br />

et al., 1993) actually advocate achiev<strong>in</strong>g this through a<br />

range of targeted activities at various spatial and temporal<br />

scales.<br />

Most of <strong>the</strong> restoration literature also po<strong>in</strong>ts towards a<br />

consensus that a ‘process-based’ approach is superior to a<br />

‘form-based’ one (Wheaton et al., 2004a). Much of <strong>the</strong><br />

form versus process debate simplifi es down to <strong>the</strong> diffi -<br />

culty and/or appropriateness <strong>in</strong> select<strong>in</strong>g an analogue or<br />

reference condition. The frequently referenced ‘Lietbilds’<br />

or target visions (Kern, 1992) and <strong>the</strong> popular Rosgen<br />

approach to restoration (Malakoff, 2004; Rosgen, 1996)<br />

both rely heavily on analogues. Jungwirth et al. (2002)<br />

suggest that at least three methods for select<strong>in</strong>g analogue<br />

or reference conditions exist:<br />

Select an exist<strong>in</strong>g reference site with ‘desirable’ conditions<br />

(location substitution).<br />

Select a historical reference condition for <strong>the</strong> site of<br />

<strong>in</strong>terest on <strong>the</strong> basis of historical analysis (time for space<br />

substitution).<br />

Create a reference condition on <strong>the</strong> basis of <strong>the</strong>oretical<br />

models (ei<strong>the</strong>r conceptual or ma<strong>the</strong>matical).<br />

In referr<strong>in</strong>g to <strong>the</strong>se analogue conditions, is <strong>the</strong> desired<br />

form or <strong>the</strong> desired process <strong>the</strong>n mimicked? This seems<br />

to be <strong>the</strong> po<strong>in</strong>t of departure for op<strong>in</strong>ions with<strong>in</strong> <strong>the</strong> restoration<br />

literature. Some argue that any mimick<strong>in</strong>g of reference<br />

conditions is a form-based approach (McDonald et al.,<br />

2004). O<strong>the</strong>rs suggest that as long as ample consideration<br />

of susta<strong>in</strong><strong>in</strong>g processes and desired functions is made, <strong>the</strong><br />

use of analogue conditions can be process based (Palmer<br />

et al., 1997; Wheaton et al., 2004a). Although exact <strong>in</strong>terpretations<br />

are <strong>the</strong>mselves uncerta<strong>in</strong> and will cont<strong>in</strong>ue to<br />

spur debate over semantics (confl ict<strong>in</strong>g evidence uncerta<strong>in</strong>ties),<br />

most concur that consideration of susta<strong>in</strong><strong>in</strong>g<br />

processes is fundamental (Wheaton et al., 2004c).<br />

Fundamental methodological disagreements arise <strong>in</strong> <strong>the</strong><br />

restoration literature with respect to passive versus active<br />

approaches to river restoration (Edmonds et al., 2003;<br />

Wissmar and Beschta, 1998). Here, active approaches are<br />

referred to as those which <strong>in</strong>volve direct structural modifi<br />

cation to <strong>the</strong> river, its fl oodpla<strong>in</strong> or <strong>in</strong>frastructure <strong>the</strong>re<strong>in</strong><br />

(e.g. channel realignment, levee removal, <strong>in</strong>stream habitat<br />

structures). By contrast, passive approaches are those that<br />

‘rely on <strong>the</strong> river to do <strong>the</strong> work’ (e.g. fl ow augmentation,


32 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Figure 3.4 Five philosophical attitudes towards uncerta<strong>in</strong>ty (The Venn diagram is meant to illustrate <strong>the</strong> overlap between contemporary<br />

attitudes towards uncerta<strong>in</strong>ty. Note that ignor<strong>in</strong>g uncerta<strong>in</strong>ty shares no overlap with contemporary attitudes towards uncerta<strong>in</strong>ty.)<br />

change <strong>in</strong> landuse, manag<strong>in</strong>g nonpo<strong>in</strong>t sources of pollution,<br />

buffer strips) (Wissmar and Beschta, 1998). Us<strong>in</strong>g a<br />

‘process-based’ approach can make <strong>in</strong>tuitive sense for<br />

passive approaches to restoration. For example, provid<strong>in</strong>g<br />

fl ow releases from a reservoir to mimic a natural hydrograph<br />

and encourage mobilisation and reorganisation of<br />

sediments, may restore <strong>the</strong> processes that ‘allow <strong>the</strong> river<br />

to do <strong>the</strong> work’ (Stanford et al., 1996; Trush et al., 2000).<br />

However, active approaches are considered favourable<br />

when natural or passive recovery may take an unacceptably<br />

long time (Montgomery and Bolton, 2003). The<br />

choice of a passive versus active approach will depend<br />

very much on <strong>the</strong> specifi c social, political, economic<br />

and environmental cont<strong>in</strong>gencies of <strong>in</strong>dividual river bas<strong>in</strong>s<br />

(Wissmar et al., 2003), as well as <strong>the</strong> extent to which<br />

<strong>in</strong>itial conditions matter (Phillips, 2002). Wheaton et al.<br />

(2004b) suggested that <strong>in</strong> some spawn<strong>in</strong>g habitat rehabilitation<br />

contexts, it may be appropriate to employ passive<br />

approaches like gravel augmentation <strong>in</strong> conjunction with<br />

active approaches like spawn<strong>in</strong>g bed enhancement to kickstart<br />

recovery. Ultimately, all <strong>the</strong>se choices are fuelled by<br />

an uncerta<strong>in</strong> conceptual understand<strong>in</strong>g of <strong>the</strong> system and<br />

logical ideas about how best to proceed with restoration.<br />

Given <strong>the</strong>se <strong>in</strong>herent uncerta<strong>in</strong>ties, adaptive management<br />

is well suited to allow practitioners and decision makers<br />

to make a decision <strong>in</strong> <strong>the</strong> face of uncerta<strong>in</strong>ty, and to adjust<br />

that decision as time and new challenges unfold (Clark,<br />

2002; Lister, 1998).<br />

7<br />

See Section 3.3.1 for <strong>the</strong> relationship between expectation and<br />

uncerta<strong>in</strong>ty.<br />

3.4 PHILOSOPHIES OF UNCERTAINTY<br />

So, is all this uncerta<strong>in</strong>ty bad? By this po<strong>in</strong>t, it should be<br />

clear that uncerta<strong>in</strong>ty <strong>in</strong> river restoration is ubiquitous.<br />

However, different segments of society view uncerta<strong>in</strong>ty <strong>in</strong><br />

very different ways, depend<strong>in</strong>g on <strong>the</strong> context (Lemons<br />

and Victor, see Chapter 1). As already mentioned, ord<strong>in</strong>ary<br />

people are quite comfortable with <strong>the</strong> uncerta<strong>in</strong>ties of<br />

life <strong>in</strong> an <strong>in</strong>tuitive and nonexplicit sense (Anderson et al.,<br />

2003; Pollack, 2003). However, uncerta<strong>in</strong>ty <strong>in</strong> policy and<br />

science, especially as reported <strong>in</strong> <strong>the</strong> media (Riebeek,<br />

2002), are very different contexts. The choice of what to do<br />

about uncerta<strong>in</strong>ty is a philosophical question. Five potential<br />

philosophical treatments of uncerta<strong>in</strong>ty are proposed <strong>in</strong><br />

Figure 3.4. Each of <strong>the</strong>se philosophies is reviewed <strong>in</strong> <strong>the</strong><br />

rema<strong>in</strong><strong>in</strong>g sections and l<strong>in</strong>ked to current attitudes with<strong>in</strong><br />

different segments of <strong>the</strong> river restoration community.<br />

3.4.1 Ignore <strong>Uncerta<strong>in</strong>ty</strong><br />

It has already been argued here that <strong>the</strong> restoration community<br />

has tended to passively ignore uncerta<strong>in</strong>ty and<br />

possible explanations as to why this may be <strong>the</strong> case proposed.<br />

For example, managers, policy and decision makers<br />

are fearful of admitt<strong>in</strong>g uncerta<strong>in</strong>ties, as this may be seen<br />

as a sign of weakness (Clark, 2002; Levy et al., 2000).<br />

Now that public support exists for river restoration, so<br />

too does <strong>the</strong> expectation7 that <strong>the</strong> problems restoration<br />

addresses are well understood. Indeed, <strong>the</strong>se problems are<br />

reasonably well understood, but numerous uncerta<strong>in</strong>ties<br />

rema<strong>in</strong>. Aside from basic, and potentially reducible,


communication uncerta<strong>in</strong>ties <strong>the</strong> signifi cance of <strong>the</strong> vast<br />

majority of uncerta<strong>in</strong>ties associated with restoration are<br />

simply not known. Admittedly, specifi c examples of<br />

uncerta<strong>in</strong>ties <strong>in</strong> restoration may <strong>in</strong>deed be <strong>in</strong>signifi cant.<br />

However, to assume <strong>in</strong>signifi cance on both ethical and<br />

technical grounds without fi rst establish<strong>in</strong>g it might ultimately<br />

backfi re on <strong>the</strong> restoration community.<br />

3.4.2 Elim<strong>in</strong>ate <strong>Uncerta<strong>in</strong>ty</strong><br />

The positivist view of <strong>the</strong> world has fuelled much scientifi<br />

c progress on <strong>the</strong> notion that uncerta<strong>in</strong>ty is bad, absolute<br />

knowledge is good, and it is necessary to strive to<br />

elim<strong>in</strong>ate uncerta<strong>in</strong>ty (Klir and Yuan, 1995; Priddy, 1999;<br />

Van Asselt and Rotmans, 2002). This fosters an unnecessarily<br />

narrow view of uncerta<strong>in</strong>ty as subsumed entirely<br />

with<strong>in</strong> <strong>the</strong> realm of science. van Asselt and Rotmans<br />

(2002) argued this view grew out of <strong>the</strong> ‘Enlightenment<br />

Period’ or ‘Age of Reason’ of <strong>the</strong> 17th and 18th centuries<br />

where science was to be ‘<strong>the</strong> provider of certa<strong>in</strong>ty.’ Fur<strong>the</strong>r<br />

to this endeavour, many scientists assumed that unique<br />

causal laws exist for all natural phenomena and ignored<br />

<strong>the</strong> possibilities of <strong>in</strong>determ<strong>in</strong>acy and equifi nality (Wilson,<br />

2001). Many physical scientists still subscribe to a ‘positivist’<br />

view (Harman, 1998), implicitly associat<strong>in</strong>g uncerta<strong>in</strong>ty<br />

with an <strong>in</strong>ability to quantify <strong>the</strong> environment, ra<strong>the</strong>r<br />

than acknowledg<strong>in</strong>g a limited understand<strong>in</strong>g about <strong>the</strong><br />

environment itself (Klir and Yuan, 1995).<br />

Whe<strong>the</strong>r specifi c types of uncerta<strong>in</strong>ty can be elim<strong>in</strong>ated<br />

depends on an <strong>in</strong>dividual’s <strong>in</strong>terpretation of semantics.<br />

Under <strong>the</strong> holistic view of uncerta<strong>in</strong>ty advocated <strong>in</strong> this<br />

chapter uncerta<strong>in</strong>ty cannot be completely elim<strong>in</strong>ated.<br />

Pollack (2003) suggests that ‘uncerta<strong>in</strong>ty is always with<br />

us and can never be fully elim<strong>in</strong>ated’. O<strong>the</strong>r authors (e.g.<br />

Knight, 1921) suggest that some types of uncerta<strong>in</strong>ty can<br />

be transformed <strong>in</strong>to related concepts (e.g. error, expectation,<br />

reliability, risk) with <strong>the</strong> help of ma<strong>the</strong>matical constructs<br />

and knowledge ga<strong>in</strong>ed from historical <strong>in</strong>ference.<br />

Through this transformation, uncerta<strong>in</strong>ty of a specifi c type<br />

(i.e. uncerta<strong>in</strong>ty for which a valid basis for classifi cation<br />

exists) <strong>in</strong> a sense might be ‘elim<strong>in</strong>ated.’ Such a transformation<br />

represents an improved understand<strong>in</strong>g of uncerta<strong>in</strong>ty<br />

but does not truly ‘elim<strong>in</strong>ate’ it.<br />

With technological progress has come <strong>the</strong> expectation<br />

of greater predictive power. Priddy (1999) suggested, ‘<strong>the</strong><br />

strictest standard of truth <strong>in</strong> science is that of predictability.’<br />

Although <strong>in</strong>tuitively no one expects prediction to be<br />

completely free of uncerta<strong>in</strong>ty, <strong>the</strong> notion that uncerta<strong>in</strong>ty<br />

can be elim<strong>in</strong>ated is latent <strong>in</strong> <strong>the</strong> ma<strong>in</strong>stream media<br />

(Riebeek, 2002). Pollack (2003) argues that scientists are<br />

accustomed to deal<strong>in</strong>g with uncerta<strong>in</strong>ty explicitly, but <strong>the</strong><br />

general public’s familiarity with uncerta<strong>in</strong>ty is implicit<br />

The Scope of Uncerta<strong>in</strong>ties <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 33<br />

and often confused. Jamieson (1996) suggests that, particularly<br />

with respect to decisions about <strong>in</strong>creased environmental<br />

protections, <strong>the</strong> ‘rhetorical role of uncerta<strong>in</strong>ty<br />

claims’ are used to suggest no action should be taken until<br />

uncerta<strong>in</strong>ty is elim<strong>in</strong>ated. Hence, it is concluded that<br />

attempts to elim<strong>in</strong>ate uncerta<strong>in</strong>ty are mislead<strong>in</strong>g and<br />

founded on ignorance of <strong>the</strong> pr<strong>in</strong>ciples of uncerta<strong>in</strong>ty.<br />

3.4.3 Reduce <strong>Uncerta<strong>in</strong>ty</strong><br />

A more pragmatic view of uncerta<strong>in</strong>ty seeks to reduce,<br />

ra<strong>the</strong>r than elim<strong>in</strong>ate, those specifi c elements that are perceived<br />

as problematic (Klir and Yuan, 1995). This approach<br />

to uncerta<strong>in</strong>ty is represented diagrammatically <strong>in</strong> Figure<br />

3.4. Notice that with regards to reduc<strong>in</strong>g uncerta<strong>in</strong>ty, <strong>the</strong><br />

key questions are, <strong>in</strong> order: can it be quantifi ed, is it signifi<br />

cant and can it be constra<strong>in</strong>ed? So long as <strong>the</strong> answer<br />

is ‘yes’ to all <strong>the</strong>se questions, uncerta<strong>in</strong>ty might be reduced.<br />

However, if <strong>the</strong> opposite is true, uncerta<strong>in</strong>ty is simply<br />

ignored. To move beyond uncerta<strong>in</strong>ty as an ambiguous<br />

buzzword that will forever plague scientists and decision<br />

makers, a broader view of uncerta<strong>in</strong>ty as <strong>in</strong>formation is<br />

appropriate (Newson and Clark, see Chapter 14).<br />

3.4.4 Cope with <strong>Uncerta<strong>in</strong>ty</strong><br />

Cop<strong>in</strong>g or liv<strong>in</strong>g with uncerta<strong>in</strong>ty represents a more proactive<br />

view of deal<strong>in</strong>g with uncerta<strong>in</strong>ty than elim<strong>in</strong>ation or<br />

reduction. This approach recognises that, regardless of<br />

<strong>the</strong> signifi cance of uncerta<strong>in</strong>ty and our ability/<strong>in</strong>ability to<br />

quantify or constra<strong>in</strong> it, we are always forced to cope with<br />

it. Especially with<strong>in</strong> <strong>the</strong> hydrologic and atmospheric<br />

modell<strong>in</strong>g literature, uncerta<strong>in</strong>ty is actively recognised<br />

and specifi c methods to cope with it are cont<strong>in</strong>ually be<strong>in</strong>g<br />

proposed (e.g. Beven, 1996a; Beven, 1996b; Osidele et al.,<br />

2003; Werritty, 2002).<br />

3.4.5 Embrace <strong>Uncerta<strong>in</strong>ty</strong><br />

Despite <strong>the</strong> advantages of efforts to cope with or reduce<br />

uncerta<strong>in</strong>ty over elim<strong>in</strong>at<strong>in</strong>g it, all <strong>the</strong> preced<strong>in</strong>g still fundamentally<br />

view uncerta<strong>in</strong>ty as negative. Several authors<br />

have departed from this view towards a more progressive<br />

view of embrac<strong>in</strong>g uncerta<strong>in</strong>ty (Johnson and Brown, 2001;<br />

Newson and Clark, see Chapter 14). One of <strong>the</strong> earlier<br />

proponents of this view appears to be Holl<strong>in</strong>g (1978), who<br />

argued:<br />

‘while efforts to reduce uncerta<strong>in</strong>ty are admirable . . . if<br />

not accompanied by an equal effort to design for<br />

uncerta<strong>in</strong>ty and obta<strong>in</strong> benefi ts from <strong>the</strong> unexpected,<br />

<strong>the</strong> best of predictive models will only lead to larger


34 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

problems aris<strong>in</strong>g more quickly and more often’ (<strong>in</strong>:<br />

Levy et al., 2000).<br />

Klir and Yuan (1995) considered uncerta<strong>in</strong>ty <strong>in</strong> modell<strong>in</strong>g<br />

as ‘an important commodity . . . , which can be traded<br />

for ga<strong>in</strong>s <strong>in</strong> <strong>the</strong> o<strong>the</strong>r essential characteristics of models.’<br />

O<strong>the</strong>rs have suggested that recognis<strong>in</strong>g that not all uncerta<strong>in</strong>ty<br />

is bad will be <strong>in</strong>creas<strong>in</strong>gly important to decision<br />

makers who are forced to make decisions <strong>in</strong> <strong>the</strong> face of<br />

uncerta<strong>in</strong>ty (Clark and Richards, 2002; Pollack, 2003).<br />

Especially <strong>in</strong> long term policy analysis (<strong>the</strong> next 20–100<br />

years) decision makers are faced with what Lempert et al.<br />

(2003) referred to as ‘deep uncerta<strong>in</strong>ty’. Johnson and<br />

Brown (2001) argued that <strong>in</strong>corporat<strong>in</strong>g uncerta<strong>in</strong>ty <strong>in</strong>to<br />

restoration design allows practitioners to consider multiple<br />

causes and hypo<strong>the</strong>sised fi xes; <strong>the</strong>reby reduc<strong>in</strong>g <strong>the</strong> potential<br />

for project failure. It has been argued here that uncerta<strong>in</strong>ty<br />

is not necessarily bad, but ignorance of it can foster<br />

unrealistic expectations. Chapman and Ward (2002) argued<br />

that uncerta<strong>in</strong>ty is not just as a risk, but also an opportunity.<br />

<strong>Uncerta<strong>in</strong>ty</strong> due to natural variability, <strong>in</strong> say fl ow regime,<br />

can be a particularly good th<strong>in</strong>g, for example by promot<strong>in</strong>g<br />

habitat heterogeneity and biodiversity (Clifford et al., see<br />

Chapter 7; Montgomery and Bolton, 2003).<br />

In Figure 3.5, <strong>the</strong> notions of embrac<strong>in</strong>g uncerta<strong>in</strong>ty are<br />

syn<strong>the</strong>sised <strong>in</strong> <strong>the</strong> context of <strong>the</strong> van Asselt (2000) typology.<br />

This approach embraces uncerta<strong>in</strong>ty as <strong>in</strong>formation<br />

and its potential for help<strong>in</strong>g avoid risks, or embrac<strong>in</strong>g<br />

unforeseen opportunities. Notice that <strong>the</strong> uncerta<strong>in</strong>ties are<br />

not treated uniformly, but <strong>in</strong>stead are segregated by source<br />

(i.e. due to limited knowledge or due to variability) and<br />

type. Anderson et al. (2003) note that environmental management<br />

problems are so diverse that a s<strong>in</strong>gle approach is<br />

unlikely to be appropriate for all. Thus, Chamberl<strong>in</strong>’s<br />

(1890) idea of multiple work<strong>in</strong>g hypo<strong>the</strong>ses is emerg<strong>in</strong>g<br />

<strong>in</strong> environmental management through advocat<strong>in</strong>g pluralistic<br />

approaches (e.g. Lempert et al., 2003; Van Asselt and<br />

Rotmans, 2002).<br />

The embrac<strong>in</strong>g uncerta<strong>in</strong>ty framework proposed here<br />

emphasises this po<strong>in</strong>t by structur<strong>in</strong>g a range of questions<br />

and possible management decisions based on <strong>the</strong> specifi c<br />

uncerta<strong>in</strong>ties at hand. In <strong>the</strong> spirit of ‘susta<strong>in</strong>able uncerta<strong>in</strong>ty’<br />

as proposed by Newson and Clark (see Chapter 14),<br />

this is not at all a rigid framework but <strong>in</strong>stead a loose and<br />

adaptive guide built around an uncerta<strong>in</strong>ty typology.<br />

Unlike <strong>the</strong> four o<strong>the</strong>r philosophical treatments of uncerta<strong>in</strong>ty,<br />

this allows <strong>the</strong> restoration scientist, practitioner or<br />

decision maker to:<br />

explore <strong>the</strong> potential signifi cance (both <strong>in</strong> terms of<br />

unforeseen consequences and welcome surprises) or<br />

<strong>in</strong>signifi cance of uncerta<strong>in</strong>ties;<br />

effectively communicate uncerta<strong>in</strong>ties;<br />

eventually make adaptive, but transparent, decisions <strong>in</strong><br />

<strong>the</strong> face of uncerta<strong>in</strong>ty.<br />

3.5 CONCLUSION<br />

In this chapter a very broad picture of uncerta<strong>in</strong>ty <strong>in</strong> river<br />

restoration and environmental management has been<br />

pa<strong>in</strong>ted. This was done to unravel <strong>the</strong> ambiguities around<br />

<strong>the</strong> notion of ‘certa<strong>in</strong>ty’ <strong>in</strong> restoration and recast uncerta<strong>in</strong>ty<br />

as useful <strong>in</strong>formation. In fact, <strong>the</strong> arguments and<br />

evidence presented challenge <strong>the</strong> view of scientifi c determ<strong>in</strong>istic<br />

‘certa<strong>in</strong>ty’ and societal beliefs that certa<strong>in</strong>ty is<br />

necessary <strong>in</strong> restoration. A typology for discrim<strong>in</strong>at<strong>in</strong>g<br />

uncerta<strong>in</strong>ty was reviewed that can be used to separate<br />

uncerta<strong>in</strong>ties that can lead to unforeseen and undesirable<br />

consequences from uncerta<strong>in</strong>ties that lead to potentially<br />

welcome surprises. Many of <strong>the</strong> uncerta<strong>in</strong>ties surround<strong>in</strong>g<br />

restoration motives, notions and approaches are most seriously<br />

manifested as communication uncerta<strong>in</strong>ties. That is,<br />

<strong>in</strong>stead of be<strong>in</strong>g expressed simply as uncerta<strong>in</strong>ties due<br />

to limited knowledge, <strong>the</strong>y are ignored and miscommunicated<br />

through <strong>the</strong> restoration process <strong>in</strong> a manner that<br />

prevents transparent decision mak<strong>in</strong>g. The signifi cance of<br />

<strong>the</strong> plethora of o<strong>the</strong>r uncerta<strong>in</strong>ties alluded to is largely<br />

situation-specifi c and, to date, unexplored.<br />

Five philosophical strategies for deal<strong>in</strong>g with uncerta<strong>in</strong>ty<br />

rang<strong>in</strong>g from <strong>the</strong> status quo of ignor<strong>in</strong>g uncerta<strong>in</strong>ty<br />

to <strong>the</strong> advocated embrac<strong>in</strong>g uncerta<strong>in</strong>ty were reviewed.<br />

Traditional scientifi c research has focused on a narrow<br />

class of uncerta<strong>in</strong>ties and adopted ‘elim<strong>in</strong>ate’ and ‘reduce’<br />

uncerta<strong>in</strong>ty philosophies. It is argued that it is unethical to<br />

assume that uncerta<strong>in</strong>ty is <strong>in</strong>signifi cant. There is an<br />

<strong>in</strong>creas<strong>in</strong>g recognition <strong>in</strong> environmental management that<br />

ethical and social dimensions are <strong>the</strong> primary drivers, with<br />

scientifi c and technical dimensions play<strong>in</strong>g a secondary<br />

role (Falkenmark and Folke, 2002; Lister, 1998) 8 . Thus, an<br />

emerg<strong>in</strong>g challenge which <strong>the</strong> restoration community is<br />

faced with is comb<strong>in</strong><strong>in</strong>g <strong>the</strong>se dimensions to ‘do <strong>the</strong> right<br />

th<strong>in</strong>g right.’ Out of <strong>the</strong> decision mak<strong>in</strong>g arena has emerged<br />

<strong>the</strong> pragmatic view of cop<strong>in</strong>g with uncerta<strong>in</strong>ty. However,<br />

from <strong>the</strong> suggestions and examples <strong>in</strong> <strong>the</strong> more general<br />

environmental management literature, it is concluded that<br />

embrac<strong>in</strong>g uncerta<strong>in</strong>ty could also help transcend <strong>the</strong> scientifi<br />

c research and decision mak<strong>in</strong>g boundaries <strong>in</strong> river<br />

restoration.<br />

8 Recall <strong>the</strong> development of notions <strong>in</strong> Section 3.3.2, and <strong>the</strong> dist<strong>in</strong>ction<br />

of Falkenmark and Folke (2002) between technical concerns<br />

(e.g. ‘do<strong>in</strong>g <strong>the</strong> th<strong>in</strong>g right’) and ethical concerns (e.g.<br />

‘do<strong>in</strong>g <strong>the</strong> right th<strong>in</strong>g’).


The Scope of Uncerta<strong>in</strong>ties <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 35<br />

Figure 3.5 Framework for embrac<strong>in</strong>g uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> decision mak<strong>in</strong>g process (This framework relies on <strong>the</strong> Van Asselt (2000)<br />

typology of uncerta<strong>in</strong>ty.)


36 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

REFERENCES<br />

Alario M, Brun M. 2001. <strong>Uncerta<strong>in</strong>ty</strong> and Controversy <strong>in</strong><br />

<strong>the</strong> Science and Ethics of Environmental Policy Mak<strong>in</strong>g.<br />

Theory and Science. 2 (1): Available on <strong>the</strong> web: \url{http://<br />

<strong>the</strong>oryandscience.icaap.org/content/vol002.001/02alariobrun.<br />

html}.<br />

Anderson JL, Hilborn RW, Lackey RT, Ludwig D. 2003. Watershed<br />

restoration: Adaptive decision mak<strong>in</strong>g <strong>in</strong> <strong>the</strong> face of<br />

uncerta<strong>in</strong>ty. In: Wissmar RC, Bisson PA, Duke M (Eds), Strategies<br />

for Restor<strong>in</strong>g <strong>River</strong> Ecosystems: Sources of Variability and<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Natural and Managed Systems. American<br />

Fisheries Society: Be<strong>the</strong>sda, Maryland; 185–201.<br />

Angermeier PL, Karr JR. 1994. Biological <strong>in</strong>tegrity versus biological<br />

diversity as policy directives: protect<strong>in</strong>g biotic resources.<br />

Bioscience 44: 690–697.<br />

Beven K. 1996a. Equifi nality and uncerta<strong>in</strong>ty <strong>in</strong> geomorphological<br />

modell<strong>in</strong>g. In: Rhoads BL, Thorn CE (Eds), The Scientifi c<br />

Nature of Geomorphology. John Wiley & Sons Ltd: Chichester;<br />

289–314.<br />

Beven K. 1996b. The limits of splitt<strong>in</strong>g: Hydrology. The Science<br />

of <strong>the</strong> Total Environment. 183: 89–97.<br />

Bohn BA, Kershner JL. 2002. Establish<strong>in</strong>g aquatic restoration<br />

priorities us<strong>in</strong>g a watershed approach. Journal of Environmental<br />

Management 64 (4): 355–363.<br />

Boon PJ. 1998. <strong>River</strong> restoration <strong>in</strong> fi ve dimensions. Aquatic Conservation:<br />

Mar<strong>in</strong>e and Freshwater Ecosystems 8 (1): 257–264.<br />

Brookes A, Downs P, Sk<strong>in</strong>ner K. 1998. <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>the</strong> eng<strong>in</strong>eer<strong>in</strong>g<br />

of wildlife habitats. Journal of <strong>the</strong> Chartered Institution<br />

of Water and Environmental Management 12 (1): 25–29.<br />

Brookes A, Sear DA. 1996. Geomorphological pr<strong>in</strong>ciples for<br />

restor<strong>in</strong>g channels. In: Brookes A, Shields FD (Eds), <strong>River</strong><br />

Channel <strong>Restoration</strong>: Guid<strong>in</strong>g Pr<strong>in</strong>ciples for Susta<strong>in</strong>able<br />

Projects. John Wiley & Sons Ltd: Chichester; 75–101.<br />

Brookes A, Shields FD. 1996. <strong>River</strong> Channel <strong>Restoration</strong>:<br />

Guid<strong>in</strong>g Pr<strong>in</strong>ciples for Susta<strong>in</strong>able Projects. John Wiley &<br />

Sons Ltd: Chichester.<br />

Brown TL, LeMay HE, Bursten BE. 1994. Chemistry: The Central<br />

Science (6th Edition). Prentice Hall: Englewood Cliffs, New<br />

Jersey.<br />

Cairns J. 2003. Numeracy and susta<strong>in</strong>ability. Ethics <strong>in</strong> Science<br />

and Environmental Politics 2003: 83–91.<br />

Chamberl<strong>in</strong> TC. 1890. The method of multiple work<strong>in</strong>g hypo<strong>the</strong>ses.<br />

Science. Repr<strong>in</strong>ted 1965: 148: 754–759.<br />

Chapman C, Ward S. 2002. <strong>Manag<strong>in</strong>g</strong> Project Risk and <strong>Uncerta<strong>in</strong>ty</strong>:<br />

A Constructively Simple Approach to Decision Mak<strong>in</strong>g.<br />

John Wiley & Sons Ltd: Chichester.<br />

Clark MJ. 2002. Deal<strong>in</strong>g with uncerta<strong>in</strong>ty: adaptive approaches<br />

to susta<strong>in</strong>able river management. Aquatic Conservation:<br />

Mar<strong>in</strong>e and Freshwater Ecosystems 12: 347–363.<br />

Clark MJ, Richards KJ. 2002. Support<strong>in</strong>g complex decisions for<br />

susta<strong>in</strong>able river management <strong>in</strong> England and Wales. Aquatic<br />

Conservation: Mar<strong>in</strong>e and Freshwater Ecosystems 12: 471–483.<br />

Coll<strong>in</strong>s JP, K<strong>in</strong>zig A, Grimm NB et al. 2000. A New Urban<br />

Ecology: Model<strong>in</strong>g human communities as <strong>in</strong>tegral parts of<br />

ecosystems poses special problems for <strong>the</strong> development and<br />

test<strong>in</strong>g of ecological <strong>the</strong>ory. American Scientist 88: 416–425.<br />

Davis MA, Slobodk<strong>in</strong> LB. 2004. The science and values of restoration<br />

ecology. <strong>Restoration</strong> Ecology 12 (1): 1–3.<br />

Doyle MW, Harbor JM, Rich CF, Spacie A. 2000. Exam<strong>in</strong><strong>in</strong>g <strong>the</strong><br />

effects of urbanization on streams us<strong>in</strong>g <strong>in</strong>dicators of geomorphic<br />

stability. Physical Geography 21 (2): 155–181.<br />

Edmonds RL, Francis RC, Mantua NJ, Petersen DL. 2003.<br />

Sources of climate variability <strong>in</strong> river ecosystems. In: Wissmar<br />

RC, Bisson PA, Duke M (Eds), Strategies for Restor<strong>in</strong>g <strong>River</strong><br />

Ecosystems: Sources of Variability and <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Natural<br />

and Managed Systems. American Fisheries Society: Be<strong>the</strong>sda,<br />

Maryland; 11–37.<br />

Ellison SLR, Rossle<strong>in</strong> M, Williams A (Eds). 2000. Quantify<strong>in</strong>g<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Analytical Measurement, Guide number 4.<br />

Eurachem/CITAC: Tedd<strong>in</strong>gton, UK.<br />

Everest FH, Sedell JR. 1984. Evaluat<strong>in</strong>g effectiveness of stream<br />

enhancement projects. In: Hassler TJ (Ed), Pacifi c Northwest<br />

Stream Habitat Management Workshop. American Fisheries<br />

Society: Humboldt State Universitiy, California; 156–164.<br />

Falkenmark M, Folke C. 2002. The ethics of socioecohydrological<br />

catchment management. Hydrology and Earth<br />

System Sciences 6 (1): 1–9.<br />

FISRWG. 1998. Stream Corridor <strong>Restoration</strong>: Pr<strong>in</strong>ciples, Processes,<br />

and Practices. GPO Item No. 0120-A: Federal Interagency<br />

Stream <strong>Restoration</strong> Work<strong>in</strong>g Group, Wash<strong>in</strong>gton, DC.<br />

Frissell CA, Liss WJ, Bayles D. 1993. An <strong>in</strong>tegrated, biophysical<br />

strategy for ecological restoration of large watersheds. In:<br />

Spangenborg NE, Potts DE (Eds), Chang<strong>in</strong>g Roles <strong>in</strong> Water<br />

Resources Management and Policy. American Water Resources<br />

Association: Be<strong>the</strong>sda, Maryland; 449–456.<br />

Gilvear DJ. 1999. Fluvial geomorphology and river eng<strong>in</strong>eer<strong>in</strong>g:<br />

future roles utiliz<strong>in</strong>g a fl uvial hydrosystems framework.<br />

Geomorphology 31: 229–245.<br />

Graf WL. 1996. Geomorphology and policy for restoration of<br />

impounded American rivers: What is ‘Natural?’ In: Rhoads BL,<br />

Thorn CE (Eds), The Scientifi c Nature of Geomorphology.<br />

John Wiley & Sons Ltd: Chichester; 221–254.<br />

Graf WL. 2001. Damage control: dams and <strong>the</strong> physical <strong>in</strong>tegrity<br />

of America’s rivers. Annals of <strong>the</strong> Association of American<br />

Geographers 91: 1–27.<br />

Hansen HO, Iversen TM. 1998. The European Centre for <strong>River</strong><br />

<strong>Restoration</strong> (ECRR). In: Hansen HO, Madsen BL (Eds), <strong>River</strong><br />

<strong>Restoration</strong> ’96 ECRR: Denmark; 73–79.<br />

Harman W. 1998. Global M<strong>in</strong>d Change: The Promise of <strong>the</strong> 21st<br />

Century (2nd Ed). Berrett–Koehler Publishers, Inc.: San<br />

Francisco, California.<br />

Hildén M. 2000. The role of <strong>in</strong>tegrat<strong>in</strong>g concepts <strong>in</strong> watershed<br />

rehabilitation. Ecosystem Health 6 (1): 39–50.<br />

Holl<strong>in</strong>g CS. 1978. Adaptive Environmental Assessment and Management.<br />

John Wiley & Sons Ltd: Chichester.<br />

Hruby T. 2003. Where is <strong>the</strong> ecology <strong>in</strong> wetland restoration?<br />

Society of Ecological <strong>Restoration</strong> Northwest and Society<br />

Wetland Scientists Jo<strong>in</strong>t Regional Conference. SERNW: Portland,<br />

Oregon.<br />

Hughes RM. 1995. Defi n<strong>in</strong>g acceptable biological status by compar<strong>in</strong>g<br />

with reference conditions. In: Davis WS, Simon RF<br />

(Eds), Biological Assessment and Decision Mak<strong>in</strong>g. Lewis<br />

Publishers: Boca Raton, Florida; 31–47.


Jamieson D. 1996. Scientifi c uncerta<strong>in</strong>ty: how do we know when<br />

to communicate research fi nd<strong>in</strong>gs to <strong>the</strong> public? The Science<br />

of <strong>the</strong> Total Environment 184: 103–107.<br />

Johnson PA, Brown ER. 2001. Incorporat<strong>in</strong>g uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong><br />

design of stream channel modifi cations. Journal of <strong>the</strong> American<br />

Water Resources Association 37 (5): 1225–1236.<br />

Johnson PA, Hey RD, Brown ER, Rosgen DL. 2002. Stream restoration<br />

<strong>in</strong> <strong>the</strong> vic<strong>in</strong>ity of bridges. Journal of <strong>the</strong> American<br />

Water Resources Association 38 (1): 55–67.<br />

Johnson PA, R<strong>in</strong>aldi M. 1997. <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>the</strong> design of stream<br />

channel restorations. In: Jr. FMH and Alsaffar A (Eds), Water<br />

for a Chang<strong>in</strong>g Global Community: The 27th Congress of <strong>the</strong><br />

International Association for Hydraulic Research. American<br />

Society of Civil Eng<strong>in</strong>eers Press: San Francisco, California;<br />

451–457.<br />

Johnson PA, R<strong>in</strong>aldi M. 1998. <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> stream channel<br />

restoration. In: Ayyub BM (Ed), <strong>Uncerta<strong>in</strong>ty</strong> Model<strong>in</strong>g and<br />

Analysis <strong>in</strong> Civil Eng<strong>in</strong>eer<strong>in</strong>g. CRC Press: London; 425–<br />

437.<br />

Jungwirth M, Muhar S, Schmutz S. 2002. Re-establish<strong>in</strong>g and<br />

assess<strong>in</strong>g ecological <strong>in</strong>tegrity <strong>in</strong> river<strong>in</strong>e landscapes. Freshwater<br />

Biology 47 (4): 867–887.<br />

Kavvas ML. 1999. On <strong>the</strong> coarse-gra<strong>in</strong><strong>in</strong>g of hydrologic processes<br />

with <strong>in</strong>creas<strong>in</strong>g scales. Journal of Hydrology 217 (3–4):<br />

191–202.<br />

Kern K. 1992. Rehabilitation of streams <strong>in</strong> South-West Germany.<br />

In: Boon PJ, Calow P, Petts G (Eds), <strong>River</strong> Conservation and<br />

Management. John Wiley & Sons Ltd: Chichester; 321–336.<br />

Klir GJ, Yuan B. 1995. Fuzzy Sets and Fuzzy Logic: Theory and<br />

Applications. Prentice Hall: Upper Saddle <strong>River</strong>, New Jersey.<br />

Knight F. 1921. Risk, <strong>Uncerta<strong>in</strong>ty</strong> and Profi t. Houghton Miffl <strong>in</strong>:<br />

New York.<br />

Koehn JD, Brierley GJ, Cant BL, Lucas AM. 2001. <strong>River</strong> <strong>Restoration</strong><br />

Framework, The National <strong>River</strong>s Consortium: Land and<br />

Water Australia, Canberra, ACT.<br />

Kondolf GM. 1995a. Five elements for effective evaluation of<br />

stream restoration. <strong>Restoration</strong> Ecology 3 (2): 133–136.<br />

Kondolf GM. 1995b. Geomorphological stream channel classifi -<br />

cation <strong>in</strong> aquatic habitat restoration – Uses and limitations.<br />

Aquatic Conservation: Mar<strong>in</strong>e and Freshwater Ecosystems<br />

5 (2): 127–141.<br />

Kondolf GM. 1996. A cross section of stream channel restoration.<br />

Journal of Soil and Water Conservation 51 (2): 119–125.<br />

Kondolf GM. 2000. Some suggested guidel<strong>in</strong>es for geomorphic<br />

aspects of anadromous salmonid habitat restoration proposals.<br />

<strong>Restoration</strong> Ecology 8 (1): 48–56.<br />

Lempert RJ, Popper SW, Bankes SC. 2003. Shap<strong>in</strong>g <strong>the</strong> Next<br />

One Hundred Years: New Methods for Quantitative, Long term<br />

Policy Analysis. The Rand Pardee Center: Santa Monica,<br />

California.<br />

Levy JK, Hipel KW, Kilgour DM. 2000. Us<strong>in</strong>g environmental<br />

<strong>in</strong>dicators to quantify <strong>the</strong> robustness of policy alternatives to<br />

uncerta<strong>in</strong>ty. Ecological Modell<strong>in</strong>g 130 (1–3): 79–86.<br />

Lew<strong>in</strong> J. 2001. Alluvial systematics. In: Maddy D, Mackl<strong>in</strong> MG,<br />

Woodard JC (Eds), <strong>River</strong> Bas<strong>in</strong> Sediment Systems: Archives of<br />

Environmental Change. A.A. Balkema Publishers: Steenwijik,<br />

The Ne<strong>the</strong>rlands; 19–41.<br />

The Scope of Uncerta<strong>in</strong>ties <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 37<br />

Lister N-ME. 1998. A systems approach to biodiversity conservation<br />

plann<strong>in</strong>g. Environmental Monitor<strong>in</strong>g and Assessment 49:<br />

123–155.<br />

Malakoff D. 2004. The river Doctor: Profi le Dave Rosgen.<br />

Science. 305 (13 August): 937–939.<br />

Marmulla G. (Ed). 2001. Dams, Fish and Fisheries. Opportunities,<br />

Challenges and Confl ict Resolution, Technical Paper no.<br />

419. FAO Fisheries Department: Rome, Italy.<br />

McDonald A, Lane SN, Chalk EA, Haycock NE. 2004. <strong>River</strong>s of<br />

dreams: on <strong>the</strong> gulf between <strong>the</strong>oretical and practical aspects<br />

of an upland river restoration. Transactions of <strong>the</strong> Institute of<br />

British Geographers 29 (3): 257–281.<br />

Merkle A, Kaupenjohann M. 2000. Derivation of ecosystemic<br />

effect <strong>in</strong>dicators – method. Ecological Modell<strong>in</strong>g 130 (1–3):<br />

39–46.<br />

Merriam–Webster. 1994. Merriam Webster’s Collegiate Dictionary.<br />

Merriam–Webster, Inc.: Spr<strong>in</strong>gfi eld, Massachusetts.<br />

Miller DE, Skidmore PB, White DJ. 2001. Channel Design, Inter-<br />

Fluve, Inc., Prepared for Wash<strong>in</strong>gton Department of Fish and<br />

Wildlife: Boseman, Montana.<br />

Montgomery DR, Bolton SM. 2003. Hydrogemorphic variability<br />

and river restoration. In: Wissmar RC, Bisson PA, Duke M<br />

(Eds), Strategies for Restor<strong>in</strong>g <strong>River</strong> Ecosystems: Sources of<br />

Variability and <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Natural and Managed Systems.<br />

American Fisheries Society: Be<strong>the</strong>sda, Maryland; 39–80.<br />

Moss T. 2004. The governance of land use <strong>in</strong> river bas<strong>in</strong>s: prospects<br />

for overcom<strong>in</strong>g problems of <strong>in</strong>stitutional <strong>in</strong>terplay with<br />

<strong>the</strong> EU Water Framework Directive. Land Use Policy 21:<br />

85–94.<br />

Mount JF. 1995. California <strong>River</strong>s and Streams: The Confl ict<br />

Between Fluvial Process and Land Use. University of California<br />

Press: Berkeley, California.<br />

NAP. 2002. Riparian Areas: Functions and Strategies for Management.<br />

National Academy Press: Wash<strong>in</strong>gton, DC.<br />

Newson MD. 2002. Geomorphological concepts and tools for<br />

susta<strong>in</strong>able river ecosystem management. Aquatic Conservation:<br />

Mar<strong>in</strong>e and Freshwater Ecosystems 12 (4): 365–379.<br />

Nijland HJ, Cals MJR (Eds). 2000. <strong>River</strong> <strong>Restoration</strong> <strong>in</strong><br />

Europe: Practical Applications. ECRR: Wagen<strong>in</strong>gen, The<br />

Ne<strong>the</strong>rlands.<br />

NRC. 1992. <strong>Restoration</strong> of Aquatic Ecosystems: Science, Technology,<br />

and Public Policy. National Academy Press: Wash<strong>in</strong>gton,<br />

DC.<br />

Osidele OO, Zeng W, Beck MB. 2003. Cop<strong>in</strong>g with uncerta<strong>in</strong>ty:<br />

A case study <strong>in</strong> sediment transport and nutrient load analysis.<br />

Journal of Water Resources Plann<strong>in</strong>g and Management-Asce<br />

129 (4): 345–355.<br />

Palmer MA, Ambrose RF, Poff NL. 1997. Ecological <strong>the</strong>ory and<br />

community restoration ecology. <strong>Restoration</strong> Ecology 5 (4):<br />

291–300.<br />

Phillips JD. 2001. Cont<strong>in</strong>gency and generalization <strong>in</strong> pedology,<br />

as exemplifi ed by texture-contrast soils. Geoderma 102 (3–4):<br />

347–370.<br />

Phillips JD. 2002. Global and local factors <strong>in</strong> earth surface<br />

systems. Ecological Modell<strong>in</strong>g 149 (3): 257–272.<br />

Pister EP. 2001. Wilderness fi sh stock<strong>in</strong>g: History and perspective.<br />

Ecosystems 4: 279–286.


38 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Pollack HN. 2003. Uncerta<strong>in</strong> Science . . . Uncerta<strong>in</strong> World. Cambridge<br />

University Press: Cambridge, UK.<br />

Power M. 1999. The Audit Society: Rituals of Verifi cation. Oxford<br />

University Press: Oxford, UK.<br />

Priddy R. 1999. Science Limited. Available on <strong>the</strong> web: http://<br />

home.no.net/rrpriddy/<strong>in</strong>dexlim.html: Oslo, Norway.<br />

Rhoads BL, Wilson D, Urban M, Herricks EE. 1999. Interaction<br />

between scientists and nonscientists <strong>in</strong> community-based<br />

watershed management: Emergence of <strong>the</strong> concept of<br />

stream naturalization. Environmental Management 24 (3):<br />

297–308.<br />

Richards K, Bras<strong>in</strong>gton J, Hughes F. 2002. Geomorphic dynamics<br />

of fl oodpla<strong>in</strong>s: ecological implications and a potential modell<strong>in</strong>g<br />

strategy. Freshwater Biology 47 (4): 559–579.<br />

Riebeek H. 2002. The perception of scientifi c uncerta<strong>in</strong>ty <strong>in</strong><br />

science news writ<strong>in</strong>g. URL: http://tc.eserver.org/13650.html<br />

Accessed on: 19 September, 2003.<br />

Rob<strong>in</strong>son CT, Tockner K, Ward JV. 2002. The fauna of dynamic<br />

river<strong>in</strong>e landscapes. Freshwater Biology 47 (4): 661–677.<br />

Roni P, Beechie TJ, Bilby RE et al. 2002. A review of stream<br />

restoration and a hierarchal strategy for prioritiz<strong>in</strong>g restoration<br />

<strong>in</strong> Pacifi c Northwest watersheds. North American Journal of<br />

Fisheries Management 22: 1–20.<br />

Rosgen D. 1996. Applied <strong>River</strong> Morphology. Wildland Hydrology:<br />

Pagosa Spr<strong>in</strong>gs, Colorado.<br />

Rotmans J, Van Asselt MBA. 2001. <strong>Uncerta<strong>in</strong>ty</strong> management<br />

<strong>in</strong> Integrated Assessment Modell<strong>in</strong>g: Towards a pluralistic<br />

approach. Environmental Monitor<strong>in</strong>g and Assessment 69 (2):<br />

101–130.<br />

RRC. 2002. Manual of <strong>River</strong> <strong>Restoration</strong> Techniques. The <strong>River</strong><br />

<strong>Restoration</strong> Centre, Silsoe, UK.<br />

Schumm SA. 1991. To Interpret <strong>the</strong> Earth: Ten Ways to be Wrong.<br />

Cambridge University Press: Cambridge, UK.<br />

Schumm SA, Lichty RW. 1965. Time, space and causality <strong>in</strong><br />

geomorphology. American Journal of Science 263 (February):<br />

110–119.<br />

Sear DA. 1994. <strong>River</strong> restoration and geomorphology. Aquatic<br />

Conservation: Mar<strong>in</strong>e and Freshwater Ecosystems 4 (2):<br />

169–177.<br />

Sear DA. 1996. The Sediment System and Channel Stability. In:<br />

Brookes A, Shields FD (Eds), <strong>River</strong> Channel <strong>Restoration</strong>:<br />

Guid<strong>in</strong>g Pr<strong>in</strong>ciples for Susta<strong>in</strong>able Projects. John Wiley &<br />

Sons Ltd: Chichester, 149–177.<br />

Shields FD. 1996. Hydraulic and hydrologic stability. In: Brookes<br />

A, Shields FD (Eds), <strong>River</strong> Channel <strong>Restoration</strong>: Guid<strong>in</strong>g<br />

Pr<strong>in</strong>ciples for Susta<strong>in</strong>able Projects. John Wiley & Sons Ltd:<br />

Chichester; 103–126.<br />

Shields FD, Copeland RR, Kl<strong>in</strong>geman PC et al. 2003. Design for<br />

stream restoration. Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g 129 (8):<br />

575–584.<br />

Smeets E, Weter<strong>in</strong>gs R. 1999. Environmental Indicators: Typology<br />

and Overview. Technical Report No 25, European Environment<br />

Agency: Copenhagen, Denmark.<br />

Soar PJ, Thorne CR. 2001. Channel <strong>Restoration</strong> Design for<br />

Meander<strong>in</strong>g <strong>River</strong>s. ERDC/CHL CR-01-1, U.S. Army Corps<br />

of Eng<strong>in</strong>eers, Eng<strong>in</strong>eer Research and Development Center,<br />

Vicksburg, Massachusetts.<br />

Stanford JA et al. 1996. A general protocol for restoration of<br />

regulated rivers. Regulated <strong>River</strong>s: Research and Management<br />

12: 391–413.<br />

Stockwell BR. 2000. The Mary <strong>River</strong> and Tributaries Rehabilitation<br />

Plan: A Review of <strong>River</strong> Processes, Fluvial Geomorphology<br />

and Ecological Concepts and Their Application to a <strong>River</strong><br />

Rehabilitation Plan for a Major South-East Queensland Catchment.<br />

Masters Dissertation, University of New England, New<br />

South Wales, Australia.<br />

Tockner K, Stanford JA. 2002. <strong>River</strong><strong>in</strong>e fl ood pla<strong>in</strong>s: present<br />

state and future trends. Environmental Conservation 29 (3):<br />

308–330.<br />

Trochim WM. 2000. The Research Methods Knowledge Base,<br />

2nd Edition. Atomic Dog Publish<strong>in</strong>g: C<strong>in</strong>c<strong>in</strong>nati, Ohio.<br />

Trush WJ, McBa<strong>in</strong> SM, Leopold LB. 2000. Attributes of an alluvial<br />

river and <strong>the</strong>ir relation to water policy and management.<br />

Proceed<strong>in</strong>gs of <strong>the</strong> National Academy of Sciences of <strong>the</strong> United<br />

States of America 97 (22): 11858–11863.<br />

Van Asselt MBA. 2000. Perspectives on <strong>Uncerta<strong>in</strong>ty</strong> and Risk:<br />

The PRIMA Approach to Decision Support. Kluwer Academic<br />

Publishers: Dordrecht, The Ne<strong>the</strong>rlands.<br />

Van Asselt MBA, Rotmans J. 2002. <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>in</strong>tegrated<br />

assessment modell<strong>in</strong>g – From positivism to pluralism. Climatic<br />

Change 54 (1–2): 75–105.<br />

Vitousek PM, Mooney HA, Lubchenco J, Melillo JM. 1997.<br />

Human dom<strong>in</strong>ation of Earth’s ecosystems. Science 277:<br />

494–499.<br />

Walker J, Diamond M, Naura M. 2002. The development of<br />

Physical Quality Objectives for rivers <strong>in</strong> England and Wales.<br />

Aquatic Conservation: Mar<strong>in</strong>e and Freshwater Ecosystems<br />

12 (4): 381–390.<br />

Walters CJ. 1997. Challenges <strong>in</strong> adaptive management of riparian<br />

and coastal ecosystems. Conservation Ecology [onl<strong>in</strong>e]: http://<br />

www.consecol.org/vol1/iss2/art1.<br />

Ward JV, Tockner K, Arscott DB, Claret C. 2002. <strong>River</strong><strong>in</strong>e landscape<br />

diversity. Freshwater Biology 47 (4): 517–539.<br />

Ward T. 1998. Risk and uncerta<strong>in</strong>ty <strong>in</strong> environmental policy<br />

evaluation. In: Dore MHI, Mount TD (Eds), Global Enviornmental<br />

Economics: Equity and <strong>the</strong> Limits of Markets. Blackwell<br />

Publishers: Oxford, UK; 116–135.<br />

Werritty A. 2002. Liv<strong>in</strong>g with uncerta<strong>in</strong>ty: climate change, river<br />

fl ows and water resource management <strong>in</strong> Scotland. The Science<br />

of The Total Environment 294 (1–3): 29–40.<br />

Wheaton JM, Pasternack GB, Merz JE. 2004a. Spawn<strong>in</strong>g habitat<br />

rehabilitation – II. Us<strong>in</strong>g hypo<strong>the</strong>sis test<strong>in</strong>g and development<br />

<strong>in</strong> design, Mokelumne <strong>River</strong>, California, USA. International<br />

Journal of <strong>River</strong> Bas<strong>in</strong> Management 2 (1): 21–37.<br />

Wheaton JM, Pasternack GB, Merz JE. 2004b. Spawn<strong>in</strong>g habitat<br />

rehabilitation – I. Conceptual approach and methods. International<br />

Journal of <strong>River</strong> Bas<strong>in</strong> Management 2 (1): 3–20.<br />

Wheaton JM, Sear DA, Darby SE, Milne JA. 2004c. The International<br />

<strong>River</strong> <strong>Restoration</strong> Survey. http://www.geog.soton.ac.<br />

uk/users/WheatonJ/<strong>Restoration</strong>Survey_Cover.asp. Accessed<br />

on: 15/06/04.<br />

Wilson DW. 2001. On <strong>the</strong> Problem of Indeterm<strong>in</strong>acy <strong>in</strong> Fluvial<br />

Geomorphology. PhD Thesis, University of Southampton,<br />

Southampton, UK.


Wissmar RC, Beschta RL. 1998. <strong>Restoration</strong> and management of<br />

riparian ecosystems: a catchment perspective. Freshwater<br />

Biology 40: 571–585.<br />

Wissmar RC, Bisson PA. 2003a. Strategies for restor<strong>in</strong>g river<br />

ecosystems: Sources of variability and uncerta<strong>in</strong>ty. In: Wissmar<br />

RC, Bisson PA, Duke M (Eds), Strategies for Restor<strong>in</strong>g <strong>River</strong><br />

Ecosystems: Sources of Variability and <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Natural<br />

and Managed Systems. American Fisheries Society: Be<strong>the</strong>sda,<br />

Maryland; 3–7.<br />

Wissmar RC, Bisson PA. 2003b. Strategies for restor<strong>in</strong>g rivers:<br />

Problems and opportunites. In: Wissmar RC, Bisson PA, Duke<br />

M (Eds), Strategies for Restor<strong>in</strong>g <strong>River</strong> Ecosystems: Sources of<br />

The Scope of Uncerta<strong>in</strong>ties <strong>in</strong> <strong>River</strong> <strong>Restoration</strong> 39<br />

Variability and <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Natural and Managed Systems.<br />

American Fisheries Society: Be<strong>the</strong>sda, Maryland; 245–262.<br />

Wissmar RC, Bisson PA (Eds). 2003c. Strategies for Restor<strong>in</strong>g<br />

<strong>River</strong> Ecosystems: Sources of Variability and <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong><br />

Natural and Managed Systems. American Fisheries Society:<br />

Be<strong>the</strong>sda, Maryland.<br />

Wissmar RC, Braatne JH, Beschta RL, Rood SB. 2003. Variability<br />

of riparian ecosystems: Implications for restoration. In:<br />

Wissmar RC, Bisson PA, Duke M (Eds), Strategies for Restor<strong>in</strong>g<br />

<strong>River</strong> Ecosystems: Sources of Variability and <strong>Uncerta<strong>in</strong>ty</strong><br />

<strong>in</strong> Natural and Managed Systems. American Fisheries Society:<br />

Be<strong>the</strong>sda, Maryland; 107–127.


SECTION II<br />

Plann<strong>in</strong>g and Design<strong>in</strong>g<br />

<strong>Restoration</strong> Projects


<strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat Edited by Stephen Darby and David Sear<br />

© 2008 John Wiley & Sons, Ltd<br />

4<br />

Plann<strong>in</strong>g <strong>River</strong> <strong>Restoration</strong> Projects:<br />

Social and Cultural Dimensions<br />

4.1 INTRODUCTION<br />

G. Mathias Kondolf 1 and Chia-N<strong>in</strong>g Yang 2<br />

1 Department of Landscape Architecture and Environmental Plann<strong>in</strong>g, University of California, USA<br />

2 Department of Landscape Architecture, California State Polytechnic University, USA<br />

Nationwide, s<strong>in</strong>ce 1990, at least US$17 billion has been<br />

spent on restoration projects, a fi gure that is an underestimate<br />

because most reported costs do not <strong>in</strong>clude staff<br />

time and many projects did not report <strong>the</strong>ir costs at all<br />

(Bernhardt et al., 2005). In many areas, river restoration<br />

has become an <strong>in</strong>dustry, with nonprofi t groups, government<br />

agencies and consult<strong>in</strong>g fi rms now depend<strong>in</strong>g upon<br />

river restoration funds to support large components of<br />

<strong>the</strong>ir budgets. For example, over four fi scal years from July<br />

2000 to June 2004, over US$100 million was disbursed<br />

by <strong>the</strong> California Department of Fish and Game to recipient<br />

groups and agencies <strong>in</strong> <strong>the</strong> Fishery <strong>Restoration</strong> Grants<br />

Program for restoration projects <strong>in</strong> coastal river bas<strong>in</strong>s<br />

(F. Sime, personal communication, December 2003),<br />

mostly to construct habitat enhancement structures <strong>in</strong><br />

salmon-bear<strong>in</strong>g rivers and streams. Elsewhere <strong>in</strong> California,<br />

stream restoration projects employ many <strong>in</strong> rural communities,<br />

<strong>in</strong>clud<strong>in</strong>g former timber cutters (Hamilton,<br />

1993). In <strong>the</strong> city of Bozeman, Montana, <strong>the</strong>re is suffi cient<br />

bus<strong>in</strong>ess to support six fi rms specializ<strong>in</strong>g <strong>in</strong> restor<strong>in</strong>g<br />

trout streams.<br />

<strong>River</strong> and stream restoration can be viewed as a contemporary<br />

phase of <strong>the</strong> environmental movement. Unlike<br />

early phases of <strong>the</strong> movement, which tended to document<br />

and draw attention to <strong>the</strong> nature and extent of environmental<br />

degradation (Carson, 1962; Ehrlich, 1968) and which<br />

<strong>the</strong>refore tended to be negative or pessimistic <strong>in</strong> tone,<br />

restor<strong>in</strong>g rivers and streams has a positive, pro-active connotation.<br />

This is especially true of streams <strong>in</strong> urban neigh-<br />

borhoods, where restoration projects can provide positive<br />

re<strong>in</strong>forcement and a sense of empowerment to local<br />

groups. In many respects, <strong>the</strong> greatest benefi ts to restor<strong>in</strong>g<br />

local urban creeks are probably <strong>the</strong> social benefi ts that<br />

accrue through <strong>the</strong> process of community build<strong>in</strong>g and <strong>the</strong><br />

public environmental education achieved.<br />

Technical specialists often assume implicitly that restoration<br />

is a technical problem, and a glance at articles<br />

published <strong>in</strong> restoration-related journals shows a preponderance<br />

of papers address<strong>in</strong>g <strong>the</strong> scientifi c aspects of<br />

project design and plann<strong>in</strong>g. However, restoration can be<br />

viewed as fundamentally a social phenomenon, as it results<br />

from a societal decision to restore some functions to a<br />

river (Eden et al., 2000). Its goals and implementation<br />

approaches can be <strong>in</strong>formed by science, but <strong>the</strong>y are<br />

essentially social <strong>in</strong> nature. The very fact that restoration<br />

has become such a widespread activity refl ects a change<br />

<strong>in</strong> public attitudes towards watercourses. The current attitudes<br />

are possible only thanks to past social <strong>in</strong>vestments<br />

<strong>in</strong> waste water treatment follow<strong>in</strong>g passage of <strong>the</strong> Clean<br />

Water Act <strong>in</strong> <strong>the</strong> United States (Wolman, 1971) and<br />

comparable legislation <strong>in</strong> o<strong>the</strong>r developed countries. The<br />

result<strong>in</strong>g improvements <strong>in</strong> water quality now make human<br />

contact with urban waters desirable and ecological restoration<br />

feasible, which was not <strong>the</strong> case <strong>in</strong> <strong>the</strong> past when <strong>the</strong>se<br />

channels were open sewers. As society and culture evolve,<br />

goals change and <strong>the</strong> realm of what is ‘feasible’ <strong>in</strong> river<br />

restoration can change dramatically, <strong>in</strong>troduc<strong>in</strong>g uncerta<strong>in</strong>ties<br />

for restoration plann<strong>in</strong>g and design – uncerta<strong>in</strong>ties<br />

that do not lend <strong>the</strong>mselves to technical eng<strong>in</strong>eer<strong>in</strong>g<br />

analysis.


44 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

This chapter considers social and cultural dimensions<br />

of river restoration, and related uncerta<strong>in</strong>ties for restoration<br />

plann<strong>in</strong>g and design. While recognized as important,<br />

<strong>the</strong>y are probably less well understood by <strong>the</strong> agencies<br />

<strong>in</strong>volved <strong>in</strong> design<strong>in</strong>g and implement<strong>in</strong>g restoration projects<br />

(FISCRWG, 1998). Systematic studies of public<br />

attitudes and expectations regard<strong>in</strong>g river restoration<br />

(Tunstall et al., 2000) have been rare <strong>in</strong> light of <strong>the</strong> enormous<br />

societal <strong>in</strong>vestment <strong>in</strong> restoration projects. There is<br />

no way this short chapter can do justice to this broad topic,<br />

but here<strong>in</strong> we attempt to raise some issues relevant to <strong>the</strong><br />

enterprise of river restoration, which we hope will be<br />

useful to scientists and practitioners <strong>in</strong> <strong>the</strong> fi eld. We fi rst<br />

briefl y review some important social aspects of river restoration:<br />

<strong>the</strong> land-use context of restoration projects,<br />

underly<strong>in</strong>g cultural preferences <strong>in</strong> river restoration design<br />

and <strong>the</strong> <strong>in</strong>creas<strong>in</strong>g importance of public participation<br />

<strong>in</strong> river management and restoration programs. We<br />

draw upon recent research to consider various human<br />

activities <strong>in</strong> urban streams and <strong>the</strong> confl icts among restoration<br />

goals of various professionals and stakeholder<br />

groups. F<strong>in</strong>ally, we present two brief case studies from<br />

nor<strong>the</strong>rn California, which illustrate social issues and<br />

attendant uncerta<strong>in</strong>ties <strong>in</strong> river restoration.<br />

4.2 OVERVIEW OF SOCIAL ASPECTS OF<br />

RIVER RESTORATION<br />

4.2.1 An Urban–Rural–Wilderness Cont<strong>in</strong>uum<br />

Appropriate goals and <strong>the</strong> solutions possible vary widely<br />

with context, from near wilderness to dense urban sett<strong>in</strong>gs<br />

(Figure 4.1). Where catchment processes are relatively<br />

unaltered and runoff and sediment load are virtually<br />

Attributes<br />

General approach<br />

Examples<br />

WILDERNESS<br />

Unaltered watershed.<br />

Channel may be altered.<br />

Can restore pre-disturbance,<br />

historical channel, by ei<strong>the</strong>r:<br />

1) lett<strong>in</strong>g river restore itself<br />

2) ‘carbon-copy’ approach<br />

Flow regime unchanged.<br />

Sediment load unchanged.<br />

No urban encroachment.<br />

Figure 4.1 An urban–wilderness cont<strong>in</strong>uum <strong>in</strong> river restoration<br />

unchanged, a restoration project can logically seek to<br />

restore pre-disturbance channel conditions, ei<strong>the</strong>r by<br />

giv<strong>in</strong>g fl oods and sediment transport <strong>the</strong> opportunity to<br />

recreate natural channel conditions (an approach often<br />

termed ‘passive restoration’) or by proactively reconstruct<strong>in</strong>g<br />

pre-disturbance channel form (<strong>the</strong> ‘carboncopy’<br />

approach of Brookes and Shields, 1996). An example<br />

would be a channel whose catchment land use has rema<strong>in</strong>ed<br />

constant, but whose form was altered by channel straighten<strong>in</strong>g<br />

or by removal of bank vegetation and consequent<br />

<strong>in</strong>stability. At this wilderness end of <strong>the</strong> cont<strong>in</strong>uum, it<br />

makes sense to ei<strong>the</strong>r let natural processes accomplish<br />

<strong>the</strong> restoration or to use <strong>the</strong> pre-disturbance channel as a<br />

template, because <strong>the</strong> processes that supported <strong>the</strong> predisturbance<br />

channel will tend to support <strong>the</strong> same channel<br />

form aga<strong>in</strong>. In ei<strong>the</strong>r case, to ma<strong>in</strong>ta<strong>in</strong> ecological values,<br />

<strong>the</strong> channel should be permitted to migrate freely and to<br />

fl ood overbank areas (Ward and Stanford, 1995).<br />

Mov<strong>in</strong>g towards <strong>the</strong> urbanized end of <strong>the</strong> cont<strong>in</strong>uum,<br />

land use change <strong>in</strong> <strong>the</strong> catchment has altered runoff<br />

and sediment load, so <strong>the</strong>re is no reason to expect predisturbance<br />

channel dimensions to be ma<strong>in</strong>ta<strong>in</strong>ed by<br />

current processes. At <strong>the</strong> extreme, urban development <strong>in</strong><br />

<strong>the</strong> catchment <strong>in</strong>creases peak fl ows such that <strong>the</strong> channel<br />

tends to <strong>in</strong>cise, which, if uncontrolled, may lead to bank<br />

collapse and channel widen<strong>in</strong>g. However, encroachment<br />

of urban development to <strong>the</strong> channel marg<strong>in</strong>s means that<br />

<strong>in</strong>cision and channel widen<strong>in</strong>g are socially unacceptable.<br />

At this urban extreme, restoration projects must be built<br />

to convey urban runoff without fl ood<strong>in</strong>g adjacent lands<br />

and to withstand <strong>in</strong>creased shear stresses of urban runoff<br />

without erosion. Here, restoration can be viewed as a form<br />

of garden<strong>in</strong>g, <strong>in</strong> which <strong>the</strong> elements are deliberately<br />

chosen and ma<strong>in</strong>ta<strong>in</strong>ed by human <strong>in</strong>put, albeit one that<br />

HIGHLY URBAN<br />

Transitional cases. Highly altered watershed and channel.<br />

Encroached banks.<br />

Can partly restore processes.<br />

Must decide what changes to accept as<br />

constra<strong>in</strong>ts, what to try to change/restore.<br />

Increase releases from dam?<br />

Reduce peak urban flow by detention?<br />

Add gravel below dams?<br />

Reduce erosion <strong>in</strong> watershed?<br />

Remove houses along bank/floodpla<strong>in</strong>?<br />

Cannot restore historical conditions.<br />

<strong>Restoration</strong><br />

as<br />

‘ garden<strong>in</strong>g:<br />

’<br />

choose elements to <strong>in</strong>clude but must account for<br />

erosive forces, altered hydrology.<br />

Social issues are important: potential to improve<br />

ecology/water quality are limited, so emphasis on<br />

community-build<strong>in</strong>g and environmental education.<br />

Flow regime altered.<br />

Sediment load altered.<br />

Urban encroachment to banks.<br />

Attributes<br />

General approach<br />

Examples


equires hard structures to resist erosive forces of urbanrunoff-augmented<br />

fl oods. In such cases, ‘naturalness’ <strong>in</strong><br />

restoration may be viewed as more an aes<strong>the</strong>tic choice<br />

than a real design approach. Such channels can be highly<br />

successful <strong>in</strong> provid<strong>in</strong>g recreational and aes<strong>the</strong>tic amenities,<br />

l<strong>in</strong>k<strong>in</strong>g communities through walk<strong>in</strong>g and bik<strong>in</strong>g<br />

trails, even provid<strong>in</strong>g for kayak<strong>in</strong>g and canoe<strong>in</strong>g. However,<br />

<strong>the</strong>y may be viewed as ‘water features’ (to borrow a term<br />

from <strong>the</strong> fi eld of landscape architecture) capable of convey<strong>in</strong>g<br />

fl oodwaters with<strong>in</strong> <strong>the</strong> channel and without erod<strong>in</strong>g<br />

banks, ra<strong>the</strong>r than large scale, dynamic ecosystems. They<br />

can still provide ecological benefi ts at a local scale, but<br />

without rebuild<strong>in</strong>g of <strong>the</strong> urban <strong>in</strong>frastructure <strong>the</strong>se waterways<br />

are unlikely to support sensitive target species at a<br />

large scale. In <strong>the</strong>se highly urban sett<strong>in</strong>gs, <strong>the</strong> ecological<br />

potential from a restoration project can rarely be comparable<br />

to that achieved <strong>in</strong> a less urban sett<strong>in</strong>g, and thus<br />

urban restoration projects may be best justifi ed by <strong>the</strong>ir<br />

potential social benefi ts as <strong>the</strong>y respond to human needs<br />

and uses.<br />

Many restoration projects can be seen to fall on a cont<strong>in</strong>uum<br />

between <strong>the</strong>se two extremes, with constra<strong>in</strong>ts, but<br />

with <strong>the</strong> potential to restore some natural processes and<br />

functions. It is <strong>in</strong> <strong>the</strong>se <strong>in</strong>termediate cases that <strong>the</strong> greatest<br />

uncerta<strong>in</strong>ties arise, as one person’s ‘constra<strong>in</strong>t’ may be<br />

ano<strong>the</strong>r’s opportunity to restore process. For example, if<br />

an upstream reservoir has elim<strong>in</strong>ated <strong>the</strong> natural magnitude<br />

and frequency of fl oods, should we accept this as<br />

a constra<strong>in</strong>t that effectively limits <strong>the</strong> degree to which<br />

natural ecosystem processes can be restored, or do we<br />

seek to alter <strong>the</strong> reservoir operation rules to more closely<br />

mimic natural fl ow patterns? Reservoir release patterns<br />

have been altered and aquatic ecological conditions<br />

improved on rivers such as <strong>the</strong> Green <strong>River</strong>, Kentucky<br />

(Postel and Ritcher, 2004), <strong>the</strong> St Mary <strong>River</strong>, Alberta<br />

(Rood and Mahoney, 2000) and Putah Creek, California<br />

(Marchetti and Moyle, 2001). Similarly, does <strong>the</strong> existence<br />

of human <strong>in</strong>frastructure or hous<strong>in</strong>g on a fl oodpla<strong>in</strong><br />

mean we cannot <strong>in</strong>undate this fl oodpla<strong>in</strong>? Or should <strong>the</strong><br />

restoration project <strong>in</strong>clude compensation for mov<strong>in</strong>g <strong>the</strong><br />

<strong>in</strong>appropriately-sited land use to higher ground, so overbank<br />

fl ood<strong>in</strong>g processes can be restored? These are social/<br />

political decisions, which can be <strong>in</strong>formed by science, but<br />

which cannot be predicted technically – add<strong>in</strong>g substantial<br />

uncerta<strong>in</strong>ty to restoration plann<strong>in</strong>g.<br />

4.2.2 Cultural Preferences <strong>in</strong> <strong>River</strong><br />

<strong>Restoration</strong> Design<br />

Unstated and often unacknowledged cultural preferences<br />

probably underlie many restoration design decisions. For<br />

example, grassy banks are preferred over shrubby or<br />

Plann<strong>in</strong>g <strong>River</strong> <strong>Restoration</strong> Projects: Social and Cultural Dimensions 45<br />

wooded banks along many restored streams <strong>in</strong> nor<strong>the</strong>rn<br />

Europe, refl ect<strong>in</strong>g <strong>the</strong> long history of pastoral land use.<br />

Similarly, open park-like landscape seems to be broadly<br />

preferred <strong>in</strong> western culture (Appleton, 1975), and residents<br />

near urban stream restoration projects <strong>in</strong> nor<strong>the</strong>rn<br />

California have compla<strong>in</strong>ed when restored streams become<br />

too ‘bushy’ and woody riparian vegetation blocks visual<br />

access to <strong>the</strong> stream bed (Purcell et al., 2002). Similarly,<br />

large woody debris <strong>in</strong> channels imparts a messy look, to<br />

which most people have a negative reaction (Piégay et al.,<br />

2005).<br />

In North America, restoration projects seek<strong>in</strong>g to create<br />

stable, symmetrically-meander<strong>in</strong>g channels have proliferated.<br />

In some cases, previously s<strong>in</strong>gle-thread channels<br />

have been reconstructed <strong>in</strong> attempts to create a more ideal,<br />

symmetrical meander<strong>in</strong>g form <strong>in</strong> <strong>the</strong> belief that <strong>the</strong>se<br />

would be more stable (Smith and Prestegaard, 2005). In<br />

o<strong>the</strong>r cases, <strong>the</strong> channels have been reconstructed with <strong>the</strong><br />

goal of convert<strong>in</strong>g braided rivers to s<strong>in</strong>gle-thread, meander<strong>in</strong>g<br />

rivers. In many cases, <strong>the</strong> streams so ‘restored’ were<br />

never s<strong>in</strong>gle-thread meander<strong>in</strong>g channels under natural<br />

conditions, and <strong>the</strong> projects can be viewed as essentially<br />

attempts to impose an idealized meander<strong>in</strong>g form onto <strong>the</strong><br />

river, as illustrated on Uvas Creek, California (Kondolf<br />

et al., 2001). Many of <strong>the</strong>se channel reconstructions have<br />

washed out with<strong>in</strong> months or years (Figure 4.2). Despite<br />

its mixed record of performance, <strong>the</strong> design approach<br />

underly<strong>in</strong>g most of <strong>the</strong>se projects – application of <strong>the</strong><br />

classifi cation scheme of Rosgen (1994) (NRC, 1992) –<br />

cont<strong>in</strong>ues to be popular among government agencies<br />

responsible for fund<strong>in</strong>g restoration projects. This is<br />

probably due to <strong>the</strong> ease with which <strong>the</strong> classifi cation<br />

scheme can be used and applied by those without academic<br />

tra<strong>in</strong><strong>in</strong>g <strong>in</strong> fl uvial geomorphology, <strong>the</strong> availability<br />

of commercial short courses teach<strong>in</strong>g users how to apply<br />

<strong>the</strong> scheme and – though largely unrealized and unacknowledged<br />

– <strong>the</strong> likelihood that <strong>the</strong> channel designs that<br />

result from apply<strong>in</strong>g <strong>the</strong> scheme satisfy a deep-seated<br />

cultural preference for stable, s<strong>in</strong>gle-thread meander<strong>in</strong>g<br />

channels.<br />

Research on human responses to landscape form suggest<br />

that subjects (at least <strong>in</strong> western culture) tend to prefer<br />

<strong>the</strong> ‘defl ected vistas’ <strong>in</strong> curved paths, rivers and valleys<br />

over straight l<strong>in</strong>es (Appleton, 1975), <strong>in</strong> part because <strong>the</strong>y<br />

elicited curiosity <strong>in</strong> subjects (Ulrich, 1983). Kaplan and<br />

Kaplan (1984) designated this landscape property as<br />

‘mystery’, convey<strong>in</strong>g <strong>the</strong> opportunity to explore and a<br />

promise to learn more with a chang<strong>in</strong>g vantage po<strong>in</strong>t as<br />

one moves more deeply <strong>in</strong>to <strong>the</strong> scene. What is probably<br />

a (near-) universal attraction to <strong>the</strong> form of meander<strong>in</strong>g<br />

channels was recognized <strong>in</strong> <strong>the</strong> 18th century by Hogarth<br />

(1753), who proposed that <strong>the</strong> ‘serpent<strong>in</strong>e’ l<strong>in</strong>e provided


46 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

a<br />

b<br />

Figure 4.2 (See also colour plate section) Uvas Creek viewed<br />

downstream from Santa Teresa Road bridge: (a) January 1996,<br />

two months after completion of <strong>the</strong> channel reconstruction<br />

project; (b) July 1997, after <strong>the</strong> constructed channel washed out<br />

<strong>in</strong> February 1996 dur<strong>in</strong>g an approximately six-year fl ow (Photo<br />

(a) courtesy of <strong>the</strong> City of Gilroy, (b) by Kondolf.)<br />

<strong>the</strong> greatest aes<strong>the</strong>tic pleasure, and more so when actively<br />

mov<strong>in</strong>g. A river mov<strong>in</strong>g through a meander<strong>in</strong>g channel<br />

thus has <strong>the</strong> elements needed for <strong>the</strong> experience of beauty<br />

under Hogarth’s <strong>the</strong>ory. This preference also found expression<br />

<strong>in</strong> <strong>the</strong> work of late 18th century English landscape<br />

designers such as Capability Brown, who built serpent<strong>in</strong>e<br />

channels on <strong>the</strong> estates of <strong>the</strong>ir wealthy clients.<br />

Despite <strong>the</strong> evidence for a landscape preference for <strong>the</strong><br />

meander<strong>in</strong>g channel form, <strong>the</strong> justifi cations for meander<strong>in</strong>g<br />

channels specifi ed <strong>in</strong> river restoration projects <strong>in</strong> North<br />

America are almost always stated <strong>in</strong> terms of bankfull<br />

discharge, width–depth ratios, meander wavelengths etc.<br />

The fundamental question as to whe<strong>the</strong>r a meander<strong>in</strong>g<br />

channel is appropriate at all is rarely addressed. Similarly,<br />

<strong>the</strong> notion that channels should be stable can be viewed<br />

as largely anthropocentric. Dynamic channels with variable<br />

fl ow regimes tend to support <strong>the</strong> greatest variety of<br />

habitats and best ecosystem function (Ward and Stanford,<br />

1995; Poff et al., 1997). Yet <strong>the</strong> meander<strong>in</strong>g channels<br />

constructed us<strong>in</strong>g <strong>the</strong> Rosgen approach have <strong>the</strong> outside<br />

of meander bends armored by root wads and boulders,<br />

with rock weirs at <strong>the</strong> crossovers to keep <strong>the</strong> ma<strong>in</strong> current<br />

away from <strong>the</strong> banks (e.g. Uvas Creek <strong>in</strong> Figure 4.2(a)).<br />

Indeed, <strong>the</strong> Rosgen scheme is used to select <strong>the</strong> ‘proper’<br />

geometry for a site, ‘proper’ mean<strong>in</strong>g it will be stable. We<br />

do not argue with <strong>the</strong> need to armor channels <strong>in</strong> dense<br />

urban areas or elsewhere when <strong>in</strong>frastructure is threatened<br />

by channel migration, but <strong>the</strong>se restoration projects typically<br />

<strong>in</strong>clude armored banks even at sites where channel<br />

migration would not threaten human works. The armor<strong>in</strong>g<br />

seems to be accepted <strong>in</strong> part because it consists of ‘natural’<br />

materials (i.e. it is not concrete) and because those <strong>in</strong>volved<br />

<strong>in</strong> fund<strong>in</strong>g and design<strong>in</strong>g <strong>the</strong>se projects hold a belief that<br />

a stable channel is preferable to an erod<strong>in</strong>g channel, even<br />

if <strong>in</strong> a rural or park sett<strong>in</strong>g.<br />

F<strong>in</strong>ally, stable, meander<strong>in</strong>g channels, fl anked by grassy<br />

banks, probably appeal to our aes<strong>the</strong>tic senses <strong>in</strong> large part<br />

because <strong>the</strong>y are ‘tidy’ landscapes. Natural riparian corridors<br />

are frequently <strong>in</strong>accessible thickets, which, while<br />

great habitat for wildlife, are unappeal<strong>in</strong>g to our western<br />

aes<strong>the</strong>tic sensibility. Nassauer (1995) demonstrated that<br />

for such ‘messy’ ecosystems to be widely accepted, we<br />

must set <strong>the</strong>m off with<strong>in</strong> a frame that conveys to <strong>the</strong><br />

viewer that <strong>the</strong> mess<strong>in</strong>ess is deliberate and not a sign of<br />

neglect. She demonstrated how ‘cues to care’ such as a<br />

neatly ma<strong>in</strong>ta<strong>in</strong>ed fence around a yard of native prairie<br />

could make <strong>the</strong> o<strong>the</strong>rwise messy bit of landscape acceptable<br />

with<strong>in</strong> <strong>the</strong> context of a suburban street.<br />

To <strong>the</strong> extent that public support for restoration is based<br />

on culturally-driven landscape preferences that are not<br />

recognized or articulated, this creates enormous uncerta<strong>in</strong>ty<br />

<strong>in</strong> river restoration projects, as public support cannot<br />

be predicted based on ‘logical’ analysis of how best to<br />

improve aquatic ecology or to manage fl oods. There is<br />

ano<strong>the</strong>r factor, which cannot be predicted by technical<br />

experts. The topic of human preference <strong>in</strong> landscape is an<br />

area of active research. Many of <strong>the</strong> fi nd<strong>in</strong>gs probably<br />

have relevance for river restoration, besides <strong>the</strong> few<br />

touched upon here.<br />

4.2.3 Public Participation and Active Stakeholders<br />

Today, public participation has become an <strong>in</strong>stitutionalized<br />

element <strong>in</strong> stream restoration. Public acceptance and<br />

support <strong>in</strong> many cases determ<strong>in</strong>es <strong>the</strong> ultimate success and<br />

susta<strong>in</strong>ability of a project. For example, provid<strong>in</strong>g public<br />

access to a restoration plan can substantially <strong>in</strong>crease<br />

public support for <strong>the</strong> plan (Bauer et al., 2002). Support<br />

for restoration is important not only <strong>in</strong> advocat<strong>in</strong>g for <strong>the</strong><br />

proposed project, but also <strong>in</strong> its stewardship after construction.<br />

Stewardship can be developed by encourag<strong>in</strong>g<br />

people to experience <strong>the</strong> restored natural areas (Ryan<br />

et al., 2002).


Increas<strong>in</strong>gly <strong>the</strong> role of stakeholders is not limited to<br />

provid<strong>in</strong>g review comments on draft documents, but<br />

to active participation <strong>in</strong> sett<strong>in</strong>g objectives and select<strong>in</strong>g<br />

implementation strategies. The success of such a collaborative<br />

plann<strong>in</strong>g process is often evaluated by whe<strong>the</strong>r or<br />

not agreement is reached among <strong>in</strong>terest groups. This<br />

approach implicitly assumes that <strong>the</strong>re is an optimal solution<br />

that satisfi es all <strong>in</strong>terests and is technically feasible.<br />

However, <strong>the</strong>re is no a priori reason to assume that this<br />

is <strong>the</strong> case, and <strong>in</strong> fact <strong>the</strong>re are good reasons to expect<br />

it frequently will not be. Accord<strong>in</strong>gly, confl icts can arise<br />

among different actors, such as between stakeholder<br />

groups, between professionals <strong>in</strong> different fi elds and<br />

between design professionals and residents, all creat<strong>in</strong>g<br />

uncerta<strong>in</strong>ties for restoration plann<strong>in</strong>g. These are discussed<br />

<strong>in</strong> more detail <strong>in</strong> <strong>the</strong> follow<strong>in</strong>g pages. Although water<br />

policy mak<strong>in</strong>g and plann<strong>in</strong>g rema<strong>in</strong>s a much contended<br />

arena <strong>in</strong> California, collaborative plann<strong>in</strong>g or policy<br />

mak<strong>in</strong>g has been documented as benefi cial not simply<br />

based on whe<strong>the</strong>r or not consent is reached among various<br />

stakeholder groups, but through <strong>the</strong> long term, <strong>in</strong>visible<br />

outcomes <strong>in</strong> terms of collective learn<strong>in</strong>g and accumulation<br />

of social, political and economic capitals (Connick and<br />

Innes, 2003).<br />

An important feature of river restoration today is <strong>the</strong><br />

proliferation of local creek groups, known <strong>in</strong> <strong>the</strong> United<br />

States as ‘Friends of’ <strong>the</strong> local creek, <strong>in</strong> <strong>the</strong> United<br />

K<strong>in</strong>gdom as river ‘Trusts’ (e.g. <strong>the</strong> Eden <strong>River</strong>s Trust). In<br />

<strong>the</strong> San Francisco Bay Area, <strong>the</strong>se groups have formed a<br />

signifi cant force <strong>in</strong> shap<strong>in</strong>g <strong>the</strong> fate of restoration projects.<br />

Friends groups not only voice <strong>the</strong>ir desires dur<strong>in</strong>g restoration<br />

plann<strong>in</strong>g, but <strong>in</strong> many cases <strong>the</strong>y have become<br />

<strong>the</strong> task force of implement<strong>in</strong>g plans and management<br />

regimes. To <strong>the</strong> restoration project designer, <strong>the</strong> potential<br />

role of local creek groups is a source of uncerta<strong>in</strong>ty. If a<br />

local group is active, it is important to work closely with<br />

it, both to improve <strong>the</strong> project design with respect to its<br />

social function<strong>in</strong>g and to improve <strong>the</strong> chances of successful<br />

implementation and susta<strong>in</strong>ability by virtue of <strong>the</strong><br />

public support a local group can often provide.<br />

There are also fundamental issues with representativeness<br />

<strong>in</strong> <strong>the</strong> stakeholder and public participation process.<br />

These processes can be drawn-out, and <strong>the</strong> long term<br />

active participants tend to be agency staff or <strong>in</strong>dustry<br />

representatives for whom participation is part of <strong>the</strong>ir job,<br />

or staff of NGOs who are often stretched th<strong>in</strong>ly amongst<br />

many such processes. To actively participate, members of<br />

<strong>the</strong> public at large must have <strong>the</strong> time and energy to devote<br />

to meet<strong>in</strong>gs over a long period (often exceed<strong>in</strong>g a year) at<br />

<strong>the</strong>ir own expense. Unless <strong>the</strong>y are strongly motivated –<br />

often by an imm<strong>in</strong>ent threat such as stopp<strong>in</strong>g a development<br />

<strong>in</strong> <strong>the</strong>ir neighborhood – few can fi nd <strong>the</strong> time to be<br />

Plann<strong>in</strong>g <strong>River</strong> <strong>Restoration</strong> Projects: Social and Cultural Dimensions 47<br />

active <strong>in</strong> <strong>the</strong> public participation process. This is refl ected<br />

by <strong>the</strong> survey results of <strong>the</strong> ‘befriended’ watersheds <strong>in</strong><br />

<strong>the</strong> San Francisco Bay Area. Neighborhoods with active<br />

‘Friends’ groups have a much higher average <strong>in</strong>come than<br />

areas that do not form creek groups (Moz<strong>in</strong>go, 2005),<br />

show<strong>in</strong>g urban stream stewardship <strong>in</strong> <strong>the</strong> United States<br />

still serves a clientele biased toward <strong>the</strong> upper and middle<br />

classes.<br />

4.3 HUMAN USES OF URBAN WATERWAYS<br />

While <strong>the</strong> habitat requirements of fi sh have been extensively<br />

studied (Reiser and Bjornn, 1979) and are used as<br />

a basis for design of restoration projects oriented towards<br />

salmon and trout (Flosi et al., 1998), <strong>the</strong> habitat requirements<br />

of humans <strong>in</strong> <strong>the</strong> stream environment, broadly<br />

construed, are less well understood. Recent research <strong>in</strong>to<br />

why certa<strong>in</strong> activities occur spontaneously at certa<strong>in</strong> parts<br />

of <strong>the</strong> stream suggests that <strong>the</strong>re are fundamental characteristics<br />

of streams that encourage, and can be designed<br />

for, recreational use. Here we review a range of human<br />

uses of stream corridors, emphasiz<strong>in</strong>g urban and suburban<br />

sett<strong>in</strong>gs.<br />

4.3.1 Camp<strong>in</strong>g by Homeless<br />

Riparian corridors have long been preferred sites for<br />

camp<strong>in</strong>g by homeless people. <strong>River</strong> corridors were sites of<br />

large camps of migratory workers and tramps <strong>in</strong> North<br />

America dur<strong>in</strong>g <strong>the</strong> depression of <strong>the</strong> 1930s, and homeless<br />

encampments are a common element along urban streams<br />

<strong>in</strong> California today, offer<strong>in</strong>g a relatively secluded refuge for<br />

<strong>the</strong> ‘down and out’. Homeless camps are often found under<br />

bridges, exploit<strong>in</strong>g <strong>the</strong> shelter from ra<strong>in</strong>, although <strong>the</strong>se<br />

sites are more accessible and thus more likely to be visited<br />

by o<strong>the</strong>rs and less private (Figure 4.3). Along Ledgewood<br />

Creek near Fairfi eld, California, a camp of fi fty residents<br />

had tents with carpeted fl oors, furniture, and batterypowered<br />

television; <strong>the</strong> residents reportedly left dur<strong>in</strong>g<br />

periodic police sweeps, only to return (Fagan, 2005).<br />

Migrants from <strong>the</strong> prov<strong>in</strong>ces of Cuba have settled along<br />

<strong>the</strong> banks of <strong>the</strong> Almendares <strong>River</strong> <strong>in</strong> Havana, form<strong>in</strong>g a<br />

squatter community known as ‘El Fangito’ (Figure 4.4).<br />

While <strong>the</strong> streets are mud, many of <strong>the</strong>se dwell<strong>in</strong>gs feature<br />

cement or tiled fl oors, furniture and television sets.<br />

Although <strong>the</strong> settlement was illegal, utilities have hooked<br />

up electrical power and water; sewage fl ows mostly<br />

through buried pipes directly to <strong>the</strong> river. The fl oodpla<strong>in</strong><br />

occupied by El Fangito is fl ooded every few years. Residents<br />

take <strong>the</strong>ir television sets and leave for higher ground<br />

when <strong>the</strong> river beg<strong>in</strong>s to rise. The management plan of <strong>the</strong><br />

Metropolitan Park of Havana (Fornes, 1994) calls for


48 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Figure 4.3 (See also colour plate section) Homeless campsite,<br />

San Pablo Creek, California (Photo by Kondolf, February<br />

2005.)<br />

Figure 4.4 (See also colour plate section) The squatter neighborhood<br />

El Fangito, Havana (Photo by Kondolf, March 2005.)<br />

mov<strong>in</strong>g <strong>the</strong>se residents from El Fangito to better, permanent<br />

hous<strong>in</strong>g and reforestation of <strong>the</strong> fl oodpla<strong>in</strong>, but an<br />

<strong>in</strong>ternational aid agency recently granted funds to build<br />

a levee around <strong>the</strong> settlement to protect it from<br />

fl oodwaters.<br />

The authors are aware of no studies focus<strong>in</strong>g on homeless<br />

use of river<strong>in</strong>e spaces, but our fi eld studies <strong>in</strong><br />

California suggest that this user group has little tolerance<br />

for o<strong>the</strong>r user groups and vice versa. We have observed<br />

that no o<strong>the</strong>r users were present near homeless camps<br />

(usually under bridges and beh<strong>in</strong>d thickets on fl oodpla<strong>in</strong>s).<br />

In Sonoma, California, children were often warned off by<br />

parents or scared away when <strong>the</strong>y accidentally <strong>in</strong>vaded<br />

homeless territory (Yang, 2004). In Japan, while homeless<br />

people also frequently reside under bridges of urban<br />

streams, <strong>the</strong> tension was not as high, as homelessness <strong>in</strong><br />

Japan is regarded ma<strong>in</strong>ly as a product of unjust <strong>in</strong>dustrial<br />

structure (Dohi, 1999), with little association with drug<br />

abuse and crime. In America, <strong>the</strong> fl ood of homelessness<br />

dur<strong>in</strong>g <strong>the</strong> past four decades is largely attributed to <strong>the</strong><br />

failure of ‘de<strong>in</strong>stitutionalization,’ a major <strong>in</strong>itiative under<br />

<strong>the</strong> Community Mental Health program that started <strong>in</strong><br />

1963 (CCHR, 2004).<br />

Occupation of river corridors by homeless can be an<br />

important source of uncerta<strong>in</strong>ty to <strong>the</strong> outcome of river<br />

restoration efforts. Use of stream corridors by homeless<br />

people has not (to <strong>the</strong> authors’ knowledge) been encouraged<br />

by designers. However, it is clearly one of <strong>the</strong> biggest<br />

uses along many urban rivers and streams. Because <strong>the</strong><br />

presence of homeless people could discourage use by<br />

o<strong>the</strong>r groups, <strong>the</strong> actual use of a restored stream corridor<br />

may be very different from that anticipated by project<br />

designers, <strong>in</strong>troduc<strong>in</strong>g uncerta<strong>in</strong>ties.<br />

4.3.2 Fish<strong>in</strong>g<br />

Fish<strong>in</strong>g is a traditional use of rivers and streams, rang<strong>in</strong>g<br />

from subsistence fi sh<strong>in</strong>g with traps and nets to purely sport<br />

fi sh<strong>in</strong>g <strong>in</strong> which <strong>the</strong> fi sh is released back to <strong>the</strong> stream.<br />

Fisheries <strong>in</strong> urban channels range widely from wild, anadromous<br />

salmonids <strong>in</strong> urban channels <strong>in</strong> <strong>the</strong> Pacifi c Northwest<br />

of North America to warm water species pulled from<br />

<strong>the</strong> polluted waters of Asian cities. Fish<strong>in</strong>g is usually well<br />

regulated by licens<strong>in</strong>g and many streams are artifi cially<br />

stocked. Fish<strong>in</strong>g is a well documented and well studied<br />

activity <strong>in</strong> rivers, a large subject well treated elsewhere<br />

and beyond <strong>the</strong> scope of this chapter. However, we po<strong>in</strong>t<br />

out that fi sh<strong>in</strong>g has long been an important activity draw<strong>in</strong>g<br />

people to rivers, similar to o<strong>the</strong>r activities we discuss<br />

below. Improv<strong>in</strong>g fi sh habitat is cited as a goal for many<br />

restoration projects and many fund<strong>in</strong>g sources are available<br />

to improve fi sheries.<br />

4.3.3 Water Sports<br />

Urban rivers (if not so polluted as to be unpleasant) have<br />

long been used for canoe<strong>in</strong>g and fl oat<strong>in</strong>g. On summer


weekends, <strong>the</strong> Chattahoochee <strong>River</strong> near Atlanta, Georgia,<br />

is packed with young people fl oat<strong>in</strong>g downstream on tire<br />

<strong>in</strong>ner tubes or rafts. Increas<strong>in</strong>gly, more active forms of<br />

kayak<strong>in</strong>g and canoe<strong>in</strong>g are be<strong>in</strong>g designed for <strong>in</strong> urban<br />

river restoration projects. For example, <strong>the</strong> steeper, upper<br />

reaches of <strong>the</strong> restored Boulder Creek <strong>in</strong> Boulder, Colorado,<br />

have been designed as a kayak course, and kayakers<br />

and canoers commonly cont<strong>in</strong>ue downstream through <strong>the</strong><br />

town.<br />

4.4 SPONTANEOUS USES OF<br />

URBAN WATERWAYS<br />

Although recreation is commonly cited as a goal of stream<br />

restoration projects, it is often treated perfunctorily compared<br />

to o<strong>the</strong>r goals such as fl ood control and habitat,<br />

except where its value can be expressed <strong>in</strong> monetary<br />

terms. In <strong>the</strong> context of cost and benefi t analysis, emphasis<br />

on recreation necessarily narrows down to <strong>the</strong> licensed,<br />

quantifi able activities such as fi sh<strong>in</strong>g and boat<strong>in</strong>g (NRC,<br />

1992). In contrast to such vacation-orientated uses, <strong>the</strong>re<br />

is a suite of more <strong>in</strong>tuitive and unplanned activities, hereby<br />

named ‘spontaneous uses,’ that <strong>in</strong>volve direct and active<br />

<strong>in</strong>teraction with <strong>the</strong> landscape, such as skipp<strong>in</strong>g rocks,<br />

catch<strong>in</strong>g frogs, collect<strong>in</strong>g nuts and swimm<strong>in</strong>g. When<br />

human uses are considered at all <strong>in</strong> urban stream restoration<br />

projects, <strong>the</strong> focus is typically on passive uses, such<br />

as trail walk<strong>in</strong>g and social ga<strong>the</strong>r<strong>in</strong>g. The orientation is<br />

also often towards adults uses only, whereas children may<br />

have very different (and strongly felt) attitudes towards<br />

stream environments (Tunstall et al., 2004; Yang, 2004).<br />

However, a grow<strong>in</strong>g literature suggests that <strong>the</strong> more<br />

<strong>in</strong>teractive activities are crucial to <strong>the</strong> form<strong>in</strong>g of environmental<br />

awareness (Chawla, 1988; Harvey, 1989; Orr,<br />

1992) and place attachment (Owens, 1988; Hester et al.,<br />

1988; Cooper-Marcus, 1992), and are benefi cial for<br />

healthy human development (Nicholson, 1971; Kaplan,<br />

1977; Cobb, 1977; Hart, 1979; Moore, 1986). Whe<strong>the</strong>r<br />

a restored channel encourages spontaneous use or not<br />

is a source of uncerta<strong>in</strong>ty to <strong>the</strong> ‘social’ success of a<br />

restoration project.<br />

To understand <strong>the</strong> specifi c habitat characteristics that<br />

permit and encourage spontaneous uses, Yang (2004)<br />

reviewed <strong>the</strong> literature to identify probable habitat characteristics<br />

encourag<strong>in</strong>g such uses and <strong>the</strong>n undertook systematic<br />

fi eld observations, especially of children, and<br />

<strong>in</strong>terviews of children and adults <strong>in</strong> fi eld areas <strong>in</strong> California<br />

and Japan. Although many of <strong>the</strong>se <strong>in</strong>teractions were<br />

engaged ma<strong>in</strong>ly by children, <strong>the</strong>y were not enjoyed by<br />

children exclusively. Adults accompany<strong>in</strong>g children, or<br />

even among a group of adults, appeared to fully enjoy<br />

such uses. From this research, we summarized <strong>the</strong> most<br />

Plann<strong>in</strong>g <strong>River</strong> <strong>Restoration</strong> Projects: Social and Cultural Dimensions 49<br />

common types of spontaneous <strong>in</strong>teraction and <strong>the</strong>ir habitat<br />

requirements.<br />

4.4.1 Quiet and Secluded Use<br />

Users who appreciate <strong>the</strong> stream environment <strong>in</strong> a transcendent<br />

way, go to <strong>the</strong> stream for a temporary escape,<br />

enjoy <strong>in</strong>timate relationships with signifi cant o<strong>the</strong>rs and<br />

those who pursue quiet read<strong>in</strong>g, th<strong>in</strong>k<strong>in</strong>g etc., are commonly<br />

attached to a specifi c base-po<strong>in</strong>t. Their territory<br />

may seem small, but <strong>the</strong> quality demands are high and<br />

specifi c. S<strong>in</strong>ce such users can stay for hours, a certa<strong>in</strong><br />

comfort level (dry seat<strong>in</strong>g, foothold and shade) is normally<br />

required. A rock, tree root, log, or a soft grassy spot<br />

by water are particularly appeal<strong>in</strong>g. Yet more than anyth<strong>in</strong>g<br />

else <strong>the</strong>y need privacy, or visual/auditory seclusion<br />

from supervision or o<strong>the</strong>r users. Lewis (1995) highlighted<br />

<strong>the</strong> value of San Leandro Creek, California, as a secret<br />

hid<strong>in</strong>g place and unsupervised play area, with many ‘fi rsttime’<br />

events of local youth. All his <strong>in</strong>terviewees who<br />

played <strong>the</strong>re appreciated this quality of nonsupervision.<br />

For this reason, <strong>the</strong>y preferred detoured or <strong>in</strong>conspicuous<br />

access and a back screen.<br />

The view toward dense foliage, open fi eld or expression<br />

of water surface, <strong>the</strong> sound of trickl<strong>in</strong>g water, <strong>the</strong> appearance<br />

of wildlife and easy access to water all tremendously<br />

enhance <strong>the</strong> value of quiet and secluded base-po<strong>in</strong>ts, as<br />

users cite <strong>the</strong>se elements as bestow<strong>in</strong>g <strong>the</strong> heal<strong>in</strong>g power<br />

of nature. Both adults and children have been found to<br />

seek out space for quiet and secluded use (Yang, 2004).<br />

Quiet and secluded base-po<strong>in</strong>ts are easily lost, not only<br />

because privacy is often lost with <strong>in</strong>creased urbanization,<br />

but also because planners and designers usually don’t<br />

design for such spots, operat<strong>in</strong>g <strong>in</strong>stead on a design model<br />

of a cheerful park for adult socializ<strong>in</strong>g and playgrounds<br />

where all children play toge<strong>the</strong>r.<br />

4.4.2 Adventures<br />

Adventure connects known to unknown parts <strong>in</strong> <strong>the</strong> landscape,<br />

expand<strong>in</strong>g cognitive and physical territory. Adventurers<br />

walk, bike, swim, leap, climb, creep and cross to<br />

‘conquer’ a new piece of landscape. A system of basepo<strong>in</strong>ts<br />

connected by diverse, usually three-dimensional<br />

paths, plays an important role <strong>in</strong> <strong>the</strong> process of expand<strong>in</strong>g<br />

territory. Dirt paths apparently possess special values to<br />

adventurers. On Marsh Creek, California, some adults<br />

favor dirt paths for aes<strong>the</strong>tic reasons, but children preferred<br />

dirt paths for <strong>the</strong> practical reasons that <strong>the</strong>y are<br />

usually avoided by adult bikers and runners, who are often<br />

impatient with children <strong>in</strong> <strong>the</strong>ir way, and <strong>the</strong> dirt path<br />

provides more <strong>in</strong>teractive features (Yang, 2004). On dirt


50 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

paths children can jump on mounds, leap <strong>in</strong>to muddy<br />

puddles, bend under low branches, or crouch to stare <strong>in</strong>to<br />

a gopher hole. The mounds, puddles, low branches or<br />

gopher holes would all be considered undesirable and<br />

elim<strong>in</strong>ated on a paved trail, but on a dirt path <strong>the</strong>y provide<br />

tempt<strong>in</strong>g <strong>in</strong>vitations for sensorial experiences.<br />

Adventurers are particularly keen to fi nd good stream<br />

cross<strong>in</strong>g po<strong>in</strong>ts. In smaller streams users search shallow<br />

and narrow spots with stepp<strong>in</strong>g rocks to set foot on or from<br />

which to build a bridge. In large rivers swimm<strong>in</strong>g across<br />

is a common game. If a rope and tree are available and <strong>the</strong><br />

stream is narrow enough, <strong>the</strong>y sw<strong>in</strong>g across <strong>the</strong> channel.<br />

Similarly, children <strong>in</strong> Turkey Brook, London, asked for<br />

ropes to sw<strong>in</strong>g on and logs to slide down (Tunstall et al.,<br />

2004). When <strong>the</strong> slope is right and <strong>the</strong> channel not too<br />

wide, bikers or skateboarders fl y across on wheels.<br />

Metal culverts are especially attractive to adventurers:<br />

it’s easy to make loud, eerie echoes <strong>in</strong> <strong>the</strong>m, <strong>the</strong>y are secret<br />

hide-outs, <strong>the</strong>y offer <strong>the</strong> allure of a connection to somewhere<br />

else and <strong>the</strong>y are perceived somehow off-limits.<br />

4.4.3 Wildlife Contact<br />

Wildlife contact can be by simple observation or active<br />

catch<strong>in</strong>g, two dist<strong>in</strong>ct modes of <strong>in</strong>teraction. Observers are<br />

usually <strong>in</strong>terested <strong>in</strong> all life forms <strong>the</strong>y see, from little bugs<br />

to big animals such as otters and raccoons. They <strong>in</strong>teract<br />

with <strong>the</strong> stream with a highly <strong>in</strong>tensive but un<strong>in</strong>trusive<br />

way. Wildlife sight<strong>in</strong>gs often occur <strong>in</strong> unexpected, uncalculated<br />

moments, produc<strong>in</strong>g a ‘wow’ experience.<br />

Catchers are more physically active and focus on certa<strong>in</strong><br />

target species, which are small enough to catch. Catch<strong>in</strong>g<br />

wildlife along creeks has traditionally provided subsistence,<br />

but <strong>in</strong> urban areas <strong>in</strong> developed nations today, catch<strong>in</strong>g<br />

is usually based upon affi nity toward <strong>the</strong> target and a<br />

sense of achievement. Fish, frogs, tadpoles, shrimps, crawdads,<br />

crabs and <strong>in</strong>sects are fasc<strong>in</strong>at<strong>in</strong>g creatures for users<br />

to match wits with. The habitats of catchers are as diverse<br />

as those of <strong>the</strong>ir target species and <strong>the</strong>ir spots correspond<br />

directly to those of <strong>the</strong>ir quarry. Children who actively<br />

catch wildlife tend to be agile and will<strong>in</strong>g to access diffi -<br />

cult sites, get wet or scratched and <strong>in</strong> general are highly<br />

adaptive to <strong>the</strong>ir environments (Figure 4.5). In Marsh<br />

Creek, crawdad hunters were often observed thriv<strong>in</strong>g at<br />

<strong>the</strong> least ‘user-friendly’ spots, such as among rugged<br />

riprap under road bridges or by grassy, muddy shores.<br />

Methods of catch<strong>in</strong>g are numerous, even for <strong>the</strong> same<br />

species. They range from <strong>the</strong> bare hand to highly elaborated<br />

means and tools. Catchers <strong>in</strong> various regions <strong>in</strong><br />

Japan and California often stored captured fi sh or crawdads<br />

temporarily <strong>in</strong> a conta<strong>in</strong>er or a little pond enclosed<br />

with sand or rocks. Most of <strong>the</strong> trapped creatures were set<br />

Figure 4.5 Crawdad from Marsh Creek caught by a child and<br />

draw<strong>in</strong>g of Marsh Creek wildlife by 4th grade child (both from<br />

Yang, 2004.)<br />

free after a short time, but some captures would become<br />

pets to be enjoyed at home until <strong>the</strong>y expired. Many catchers<br />

have learned from experiences which animals ‘work<br />

better’ as pets (Yang, 2004).<br />

Profi cient catchers and observers are often knowledgeable;<br />

<strong>the</strong>y can usually identify many species and know<br />

when and where to fi nd <strong>the</strong>m. Observers and catchers have<br />

similar habitat requirements: <strong>the</strong> environment needs to<br />

support a suffi ciently high density of wildlife and a mean<strong>in</strong>gful<br />

human/wildlife <strong>in</strong>terface. Though <strong>the</strong> former is a<br />

widely claimed goal <strong>in</strong> restoration and greenway projects,<br />

<strong>the</strong> latter is usually discouraged. For spontaneous users, a<br />

mean<strong>in</strong>gful wildlife/human <strong>in</strong>terface provides plenty of<br />

chances for close-up observation and hands-on catch<strong>in</strong>g,<br />

without <strong>the</strong> need of specialized equipment beyond that<br />

which can be made at home or obta<strong>in</strong>ed from a grocery<br />

store. Examples of such <strong>in</strong>terfaces are water edges framed<br />

by vegetation or porous structures where different species<br />

hide, or shallow water reaches adjacent to gravel bars<br />

where fry of amphibians and fi sh hatch. It is important that<br />

water edges designed to susta<strong>in</strong> a dense wildlife popula-


tion also rema<strong>in</strong> accessible to users, except <strong>in</strong> cases (rare<br />

<strong>in</strong> urban areas) where protected species need to be isolated<br />

from human harassment. When physical access is not feasible,<br />

visual access can be provided from <strong>the</strong> bank, bridges<br />

etc.<br />

4.4.4 Manipulat<strong>in</strong>g <strong>the</strong> Environment<br />

The value of creeks and rivers for spontaneous use depends<br />

largely on <strong>the</strong>ir provision of loose parts – elements that<br />

can be easily manipulated <strong>in</strong> <strong>the</strong> environment (Nicholson,<br />

1971). At least three categories of common uses rely on<br />

contact with rocks, plants, junk and o<strong>the</strong>r k<strong>in</strong>ds of loose<br />

parts <strong>in</strong> stream environments: collect<strong>in</strong>g, build<strong>in</strong>g and<br />

clever craft:<br />

• Collect<strong>in</strong>g allows one to discern treasures from <strong>the</strong> basically<br />

chaotic stream environment. Once purposely rummaged<br />

or fortuitously encountered, stones and o<strong>the</strong>r<br />

elements from <strong>the</strong> bed, plant parts and junk recovered<br />

from banks may be used <strong>in</strong> drama play, build<strong>in</strong>g, or <strong>in</strong><br />

displays <strong>in</strong> <strong>the</strong> collector’s yard or room. Gravel bars are<br />

prized sources of stones for collect<strong>in</strong>g and clay banks<br />

provide material for handcraft, mud ball fi ghts and gray<br />

make-up.<br />

• Build<strong>in</strong>g projects, whe<strong>the</strong>r big (e.g. tree houses, huts,<br />

bases, dams, bridges, ponds) or small (arrang<strong>in</strong>g rocks<br />

and sticks) are rooted <strong>in</strong> an <strong>in</strong>nate attempt to create an<br />

impact on <strong>the</strong> landscape (Figure 4.6). Through build<strong>in</strong>g,<br />

users claim <strong>the</strong>ir ownership and adapt <strong>the</strong> stream to<br />

<strong>the</strong>mselves. The result of spontaneous build<strong>in</strong>g usually<br />

is not durable enough to survive fl oods and o<strong>the</strong>r natural<br />

processes. Build<strong>in</strong>g may have practical purposes, but <strong>the</strong><br />

process is all-important: many children build, destroy<br />

and rebuild.<br />

Figure 4.6 12-year-old child’s stove <strong>in</strong> drama house by Marsh<br />

Creek (from Yang, 2004.)<br />

Plann<strong>in</strong>g <strong>River</strong> <strong>Restoration</strong> Projects: Social and Cultural Dimensions 51<br />

• Clever crafts, <strong>the</strong> skillful manipulation of materials<br />

found <strong>in</strong> stream environments (Yang, 2004), is usually<br />

quite precise <strong>in</strong> terms of materials and surround<strong>in</strong>gs.<br />

For example, to skip a rock (a trans-culturally popular<br />

trick), one needs a gravel bar conta<strong>in</strong><strong>in</strong>g platy stones<br />

with <strong>in</strong>termediate axes usually between about 30 and<br />

70 mm, and a fl at pool allow<strong>in</strong>g satisfactory skips. Along<br />

Sonoma Creek <strong>in</strong> California, adults applied red algae to<br />

sk<strong>in</strong> rash and children made fl utes with deer grass.<br />

Along Kure <strong>River</strong> <strong>in</strong> Japan, smashed mugwort was used<br />

to heal scratches and defog goggles, while foxtail stalks<br />

were made <strong>in</strong>to knots to trap frogs and shrimp.<br />

4.4.5 Wad<strong>in</strong>g and Paddl<strong>in</strong>g<br />

Small children and o<strong>the</strong>r users who don’t want to get very<br />

wet will wade <strong>in</strong> waters shallower than 0.5 m, with currents<br />

20 cm/s or less, such as shallow marg<strong>in</strong>s or backwaters.<br />

The range of paddl<strong>in</strong>g by children is usually only a<br />

few meters from <strong>the</strong> water edge and <strong>the</strong> dry spot. In large<br />

streams, some h<strong>in</strong>ts of boundary around a smaller space<br />

(e.g. a cover or re-entrant <strong>in</strong> <strong>the</strong> bank or offshore bar) are<br />

needed to overcome <strong>the</strong> uneas<strong>in</strong>ess <strong>in</strong>duced by an unlimited<br />

expanse of water. Paddlers prefer gently slop<strong>in</strong>g<br />

access to water (ra<strong>the</strong>r than grassy or upright banks) and<br />

sandy or clay bottoms (which provide comfortable footholds)<br />

(Yang, 2004).<br />

4.4.6 Swimm<strong>in</strong>g, Flush<strong>in</strong>g and Div<strong>in</strong>g<br />

Swimm<strong>in</strong>g occurs mostly <strong>in</strong> pool reaches more than 0.5 m<br />

deep, with velocities under 0.5 m/s and with gentle and<br />

gradual water edges at bars or ‘ledge’ banks protected<br />

by tree roots as entry po<strong>in</strong>ts. In large or swift rivers,<br />

swimmers also require ‘stopover bases’ (island bars,<br />

bridge piers etc.) at which to rest. Warm surfaces such as<br />

big rocks, pebble beach, concrete blocks, asphalt roads<br />

etc. are valuable dry spots. Flush<strong>in</strong>g makes clever use of<br />

locally concentrated fl ow (>0.5 m/s) and variations <strong>in</strong> bed<br />

form. Most commonly, fl ush<strong>in</strong>g is done <strong>in</strong> riffl es: <strong>the</strong><br />

fl usher would start at <strong>the</strong> end of <strong>the</strong> pool where <strong>the</strong> speed<br />

starts to pick up, allow<strong>in</strong>g his body to be carried by <strong>the</strong><br />

accelerat<strong>in</strong>g current downstream to be caught at <strong>the</strong> crest<br />

of riffl e (if shallow) or carried through <strong>the</strong> riffl e (if deeper)<br />

to <strong>the</strong> next pool. Hard structures <strong>in</strong> <strong>the</strong> streams such as<br />

bridge piers can also form concentrated currents for<br />

fl ush<strong>in</strong>g.<br />

Div<strong>in</strong>g <strong>in</strong> larger streams and rivers with deep pools<br />

is popular on hot days, offer<strong>in</strong>g <strong>the</strong> thrill of a free fall and<br />

<strong>the</strong> sudden imp<strong>in</strong>gement of cool water on <strong>the</strong> body. The


52 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

div<strong>in</strong>g height is limited by pool depth and <strong>the</strong> diver’s skill<br />

and nerve. We observed children div<strong>in</strong>g from a 15-m high<br />

treetop <strong>in</strong> <strong>the</strong> Kagami <strong>River</strong>, Japan (Figure 4.7). In stream<br />

environments, a pool deeper than two meters is rare and<br />

considered enough for moderate-height div<strong>in</strong>g. A good<br />

div<strong>in</strong>g spot also has a land<strong>in</strong>g spot with a gentler water<br />

edge, a path to connect land<strong>in</strong>g and launch spots <strong>in</strong>to a loop<br />

and, ideally, choices for different skill levels. Rock outcrops<br />

(adjacent to deep pools because <strong>the</strong>y <strong>in</strong>duce scour at<br />

high fl ow) provide steady footholds for <strong>the</strong> launch po<strong>in</strong>t<br />

and outcrops with complex form and multiple take-off<br />

po<strong>in</strong>ts for different dive heights allow divers to practice<br />

and build <strong>the</strong>ir courage and skills gradually. Div<strong>in</strong>g from<br />

<strong>the</strong> trees, one experiences <strong>the</strong> thrill of shak<strong>in</strong>g footholds;<br />

div<strong>in</strong>g from a rope sw<strong>in</strong>g, one challenges <strong>the</strong> arm strength,<br />

body balance and <strong>the</strong> tim<strong>in</strong>g to let go; div<strong>in</strong>g at concrete<br />

levees, one needs to leap forward to avoid <strong>the</strong> concrete<br />

foundation jett<strong>in</strong>g out beneath <strong>the</strong> mean water level (Yang,<br />

2004). For ‘thrill<strong>in</strong>g’ water contacts such as fl ush<strong>in</strong>g and<br />

div<strong>in</strong>g, routes that connect back to <strong>the</strong> set-<strong>in</strong> po<strong>in</strong>ts are<br />

<strong>in</strong>dispensable to support <strong>the</strong>ir repetitive characteristics.<br />

Figure 4.7 Children div<strong>in</strong>g <strong>in</strong>to <strong>the</strong> Kagami <strong>River</strong>, Japan (from<br />

Yang, 2004.)<br />

4.5 CONFLICTS AMONG MULTIPLE GOALS<br />

AND OBJECTIVES<br />

Everybody wants more nature but <strong>the</strong>re has been persistent<br />

confusion about <strong>the</strong> mean<strong>in</strong>g of ‘restoration’. The controversy<br />

over <strong>the</strong> term refl ects disagreements over goals, even<br />

when considered only with<strong>in</strong> <strong>the</strong> physical science realm.<br />

As causative agents, humans constantly change and control<br />

nature to ‘help’ it, lead<strong>in</strong>g to fundamental questions about<br />

goals. Given that nature is <strong>in</strong> constant fl ux and <strong>the</strong>re is no<br />

s<strong>in</strong>gle correct condition (Hull and Robertson, 2000), <strong>the</strong><br />

choice of desired end state for restoration will perforce<br />

<strong>in</strong>volve societal priorities. Because river restoration projects<br />

are now commonly undertaken with <strong>the</strong> <strong>in</strong>volvement<br />

of multiple professionals and stakeholders, and because<br />

all <strong>the</strong> goals and objectives are essentially value-driven,<br />

three types of confl icts often arise <strong>in</strong> project implementation,<br />

creat<strong>in</strong>g uncerta<strong>in</strong>ties for <strong>the</strong> course of river restoration<br />

plann<strong>in</strong>g and implementation.<br />

4.5.1 Confl icts among Professionals<br />

Eng<strong>in</strong>eers, fl uvial geomorphologists, ecologists and landscape<br />

architects are tra<strong>in</strong>ed to see <strong>the</strong> stream differently<br />

(Figure 4.8). In <strong>the</strong> past, <strong>the</strong>y have all shaped or reshaped<br />

streams with <strong>the</strong>ir particular value systems and discipl<strong>in</strong>ary<br />

tools. Eng<strong>in</strong>eer<strong>in</strong>g has been <strong>the</strong> s<strong>in</strong>gle most powerful<br />

profession <strong>in</strong> past stream transformation, alter<strong>in</strong>g rivers<br />

for fl ood control, water supply and navigation. Hydraulic<br />

eng<strong>in</strong>eers model fl ows under conditions where <strong>the</strong> variables<br />

are controlled, usually approximat<strong>in</strong>g channel shapes<br />

as simpler geometric entities. Deviations from clear water<br />

and Euclidean channel shapes are treated with adjustments<br />

<strong>in</strong> formulas. Fluvial geomorphologists tend to approach<br />

problems at larger scales and over longer periods. The<br />

eng<strong>in</strong>eer or manager may pose a question such as, ‘What<br />

k<strong>in</strong>d of bank protection should we use along this reach<br />

of stream?’ The fl uvial geomorphologist will tend to ask<br />

why <strong>the</strong> bank is erod<strong>in</strong>g <strong>in</strong> <strong>the</strong> fi rst place, whe<strong>the</strong>r it<br />

is simply part of <strong>the</strong> natural channel migration process or<br />

a result of changes <strong>in</strong> <strong>the</strong> catchment upstream. Especially<br />

<strong>in</strong> <strong>the</strong> latter case, it is likely that plac<strong>in</strong>g bank protection<br />

will not ‘solve’ <strong>the</strong> problem but will <strong>in</strong>duce problems<br />

elsewhere.<br />

Ecologists tend to view streams as organic compounds<br />

of habitats. They perceive fi ne details of leaf litter and its<br />

decompostion, moss on boulders, food cha<strong>in</strong>s, and cycles<br />

of nitrogen, carbon etc. Traditional biologists may see<br />

unspoiled natural process as <strong>the</strong> reference condition<br />

aga<strong>in</strong>st which to measure degradation and a return to that<br />

condition as a restoration goal. Human activities are<br />

viewed as ‘impact.’ Add<strong>in</strong>g a spatial structural perspective,


<strong>the</strong>y view streams as ‘corridors,’ a crucial element <strong>in</strong> landscape<br />

to allow movement of species and <strong>the</strong>refore to ma<strong>in</strong>ta<strong>in</strong><br />

biodiversity and long term genetic diversity. These<br />

professionals have long realized that to ma<strong>in</strong>ta<strong>in</strong> a healthy<br />

ecosystem, a river needs fl oods (Poff et al., 1997; Junk<br />

et al., 1989). Developers and fl ood control agencies often<br />

resist los<strong>in</strong>g developable lands to fl ood <strong>in</strong>undation and,<br />

through <strong>the</strong>ir <strong>in</strong>fl uence on <strong>the</strong> political process, typically<br />

succeed <strong>in</strong> implement<strong>in</strong>g fl ood control measures such as<br />

dams or levees that permit <strong>the</strong>m to build on fl oodpla<strong>in</strong>s.<br />

Similarly, natural channel migration is an important<br />

process to create diverse habitats, but riverside developments<br />

are threatened by bank erosion, result<strong>in</strong>g <strong>in</strong> pressure<br />

to stabilize <strong>the</strong> river bank with hard structures. As a result,<br />

although ecologists and environmental scientists have<br />

Plann<strong>in</strong>g <strong>River</strong> <strong>Restoration</strong> Projects: Social and Cultural Dimensions 53<br />

Figure 4.8 Different perceptions and attitudes towards rivers by eng<strong>in</strong>eers, biologists etc. (Source: Hough (1990), adapted from<br />

draw<strong>in</strong>gs orig<strong>in</strong>ally prepared by Newbury (unpublished data)).<br />

been <strong>in</strong>stitutionalized <strong>in</strong>to <strong>the</strong> plann<strong>in</strong>g process s<strong>in</strong>ce <strong>the</strong><br />

1960s, <strong>the</strong>y often rema<strong>in</strong> ‘second-class citizens’ <strong>in</strong> affect<strong>in</strong>g<br />

<strong>the</strong> design of urban stream channels (Riley, 1998).<br />

A traditional tenet of landscape architects is to view<br />

landscape <strong>in</strong> abstract, formal, aes<strong>the</strong>tic terms: forms,<br />

l<strong>in</strong>es, colors, textures and <strong>the</strong>ir <strong>in</strong>ter-relationships (Daniel<br />

and V<strong>in</strong><strong>in</strong>g, 1983). Although visual aes<strong>the</strong>tics are usually<br />

<strong>the</strong> paramount ‘public’ goal, designers also emphasize <strong>the</strong><br />

cultural and historical signifi cance of urban streams, as<br />

well as <strong>the</strong> user’s experiences. Some landscape architects<br />

are well tra<strong>in</strong>ed ecologically and effectively <strong>in</strong>tegrate ecological<br />

considerations <strong>in</strong> <strong>the</strong>ir designs; some are <strong>in</strong>volved<br />

<strong>in</strong> successful efforts to redevelop urban waterfronts to<br />

revitalize downtown economies, attract tourists and<br />

provide recreation opportunities for urban residents (Otto


54 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

et al., 2004). These projects provide chances to enhance<br />

physical and visual connection with streams by plac<strong>in</strong>g<br />

walkways along <strong>the</strong>m, or by promot<strong>in</strong>g vistas and fac<strong>in</strong>g<br />

commercial fronts to <strong>the</strong> streams (Jones and Battaglia,<br />

1989). Landscape architects also transform fl oodpla<strong>in</strong>s to<br />

open spaces to accommodate civic activities such as<br />

exhibits, concerts, fairs or sports. The design of <strong>the</strong>se open<br />

spaces, however, often takes its model from pastoral parks<br />

or architectural plazas and while <strong>the</strong>y may provide effective<br />

urban spaces, <strong>the</strong> purported uses and <strong>the</strong> design<br />

schemes are often <strong>in</strong> confl ict with <strong>the</strong> chaotic character of<br />

fl oods and organic quality of riparian habitats.<br />

4.5.2 Confl icts Among Stakeholder Groups<br />

Humans use rivers for many different purposes, so it<br />

should come as no surprise that human expectations and<br />

<strong>the</strong> demands of rivers often confl ict. The same confl icts<br />

that manifest <strong>the</strong>mselves <strong>in</strong> management of exist<strong>in</strong>g river<br />

channels tend to emerge when river restoration projects<br />

are conceived and scoped.<br />

Some of <strong>the</strong> best documented such confl icts are <strong>in</strong> <strong>the</strong><br />

Colorado <strong>River</strong> below Glen Canyon Dam, where cold<br />

water releases have allowed a ra<strong>in</strong>bow trout (Oncorhynchus<br />

mykiss) fi shery, highly valued by anglers, to become<br />

established. The ra<strong>in</strong>bow trout are exotic to <strong>the</strong> river and<br />

would not have survived <strong>the</strong> high temperatures and high<br />

suspended sediment loads characteristic of <strong>the</strong> pre-dam<br />

river. The post-dam river is now unfavorable to native fi sh,<br />

such as humpback chub (Cila cypha) and razorback sucker<br />

(Xyrauchen texanus). The native fi sh are ugly and undesirable<br />

as sport fi sh, but <strong>the</strong>y are native to <strong>the</strong> river and <strong>the</strong>ir<br />

numbers have dw<strong>in</strong>dled such that several species are now<br />

listed as threatened or endangered (Schmidt et al., 1998).<br />

Where <strong>the</strong> exotic trout and native fi sh coexist, <strong>the</strong> trout<br />

may prey on <strong>the</strong> natives. Proposed actions to improve<br />

conditions for <strong>the</strong> native species have encountered resistance<br />

from trout fi sh<strong>in</strong>g groups. It is well established that<br />

<strong>the</strong> reduction <strong>in</strong> high fl ows effected by Glen Canyon Dam<br />

has had numerous ecological effects on <strong>the</strong> reach downstream<br />

and thus deliberate high fl ow releases are planned<br />

<strong>in</strong> efforts to restore <strong>the</strong> reach. The fi rst such release, a<br />

much-publicized fl ow of 1300 m 3 s −1 <strong>in</strong> 1996, was only<br />

about one-third of <strong>the</strong> average annual pre-dam high fl ow.<br />

The fl ow was limited to avoid <strong>in</strong>undat<strong>in</strong>g a rare snail that<br />

had extended its range down <strong>the</strong> canyon walls dur<strong>in</strong>g <strong>the</strong><br />

post-dam period (Marzolf et al., 1998). Thus, restoration<br />

of a dynamic fl ow regime (with attendant benefi ts for <strong>the</strong><br />

river ecosystem) was perceived to confl ict with protection<br />

of <strong>the</strong> rare snail. A similar confl ict among user groups is<br />

on <strong>the</strong> North Fork Fea<strong>the</strong>r <strong>River</strong>, California, where high<br />

fl ows released periodically to provide fl ows for rafters<br />

have scoured benthic macro<strong>in</strong>vertebrates, wash<strong>in</strong>g <strong>the</strong>se<br />

and o<strong>the</strong>r organisms downstream (Garcia and Associates,<br />

2005).<br />

4.5.3 Confl icts Between Professionals and<br />

Local Groups<br />

Perhaps <strong>the</strong> best documented example of a professional–<br />

local group confl ict <strong>in</strong>volves terrestrial habitat restoration,<br />

<strong>the</strong> Chicago prairie restoration controversy. Efforts to<br />

restore 7000 acres of <strong>the</strong> DuPage County forest reserves<br />

<strong>in</strong> <strong>the</strong> Chicago metropolitan area back to <strong>the</strong> historical oak<br />

savanna and tallgrass prairie condition were attacked by<br />

local groups and residents who opposed remov<strong>in</strong>g trees<br />

and brush. Ryan (2000) concluded that <strong>the</strong> discrepancy<br />

between restoration planners and neighborhood users<br />

stemmed from differences <strong>in</strong> attachment. While both<br />

groups were attached to nature, <strong>the</strong>ir attachment can be<br />

diverse and contradictory. Scientists and volunteers are<br />

attached to a particular type of orig<strong>in</strong>al landscape, which<br />

is established through environmental criteria such as biodiversity<br />

and system <strong>in</strong>tegrity. Such attachment is not<br />

bound to a place – <strong>the</strong> same habitat image can be reproduced<br />

elsewhere and still be satisfactory. On <strong>the</strong> o<strong>the</strong>r<br />

hand, <strong>the</strong> attachment of local residents is <strong>in</strong>tertw<strong>in</strong>ed <strong>in</strong><br />

locale and context. Individual trees, albeit non-native, bear<br />

an identity <strong>in</strong> terms of furnish<strong>in</strong>g <strong>the</strong> spot for children to<br />

play tag or a seat for quietness or fram<strong>in</strong>g a magnifi cent<br />

view toward sundown. In o<strong>the</strong>r words, <strong>the</strong> attachment of<br />

local residents is composed of life memories.<br />

A different confl ict between professionals and a local<br />

group occurred <strong>in</strong> <strong>the</strong> nor<strong>the</strong>rn Sierra Nevada of California<br />

<strong>in</strong> <strong>the</strong> early 1990s. In plann<strong>in</strong>g a restoration project on<br />

Jamison Creek <strong>in</strong> Plumas-Eureka State Park (<strong>the</strong> Park), a<br />

local nonprofi t group active <strong>in</strong> implement<strong>in</strong>g stream restoration<br />

projects (but without expertise <strong>in</strong> fl uvial geomorphology)<br />

challenged <strong>the</strong> effective discharge analysis<br />

conducted by a university team, contend<strong>in</strong>g that <strong>the</strong> bankfull<br />

discharge was only about one-third that computed by<br />

<strong>the</strong> university team. Despite a thorough and well documented<br />

scientifi c report support<strong>in</strong>g <strong>the</strong> university team’s<br />

analysis, <strong>the</strong> Park rejected <strong>the</strong> analysis and sided with <strong>the</strong><br />

local nonprofi t group, stat<strong>in</strong>g that it preferred <strong>the</strong> smaller<br />

design discharge because ‘a smaller channel is better for<br />

fi sh habitat’. The Park also cited that fact that <strong>the</strong> local<br />

Coord<strong>in</strong>ated Resource Management Program group (composed<br />

of local agency staff, landowners and staff of <strong>the</strong><br />

nonprofi t group (none of whom possessed expertise <strong>in</strong><br />

fl uvial geomorphology) had voted <strong>in</strong> favor of <strong>the</strong> lower<br />

design discharge.<br />

The notion that one can arbitrarily choose a design discharge<br />

and build a stream channel to smaller dimensions


is a fasc<strong>in</strong>at<strong>in</strong>g one, but not one supported by geomorphic<br />

science. Likewise, <strong>the</strong> notion that scientifi c questions<br />

should be put to a vote by a group without expertise <strong>in</strong> <strong>the</strong><br />

fi eld raises questions about <strong>the</strong> role of science <strong>in</strong> such a<br />

restoration design process. Ultimately, <strong>the</strong> Park had <strong>the</strong><br />

authority and responsibility to design and construct <strong>the</strong><br />

channel as it saw fi t. The university team withdrew from<br />

<strong>the</strong> project and <strong>the</strong> local nonprofi t group proceeded to<br />

design and build a channel reconstruction <strong>in</strong> 1995. The<br />

project was damaged by <strong>the</strong> high fl ows of 1996, repaired,<br />

and <strong>the</strong>n completely washed out by high fl ows <strong>in</strong> 1997. In<br />

2000, <strong>the</strong> Park sent out a call for proposals to reconstruct<br />

<strong>the</strong> channel once aga<strong>in</strong>.<br />

4.6 CASE STUDIES<br />

4.6.1 Baxter Creek, El Cerrito<br />

Baxter Creek dra<strong>in</strong>s an 11-km2 urban area of El Cerrito,<br />

California, debouch<strong>in</strong>g <strong>in</strong>to San Francisco Bay at Richmond<br />

(Figure 4.9). In 1997, <strong>the</strong> City of El Cerrito replaced<br />

a 70-m reach of fail<strong>in</strong>g culvert (<strong>in</strong> a small neighborhood<br />

park) with an open channel (Figure 4.10). The open<br />

channel was stabilized with a series of boulder weirs<br />

Pacific<br />

Ocean<br />

San<br />

Francisco<br />

San<br />

Pablo<br />

Bay<br />

San<br />

Francisco<br />

Bay<br />

Plann<strong>in</strong>g <strong>River</strong> <strong>Restoration</strong> Projects: Social and Cultural Dimensions 55<br />

Suisun Bay<br />

Richmond<br />

a b<br />

Brentwood<br />

Berkeley<br />

San Jose<br />

Figure 4.9 Location map, Baxter (a) and Marsh Creeks (b),<br />

Contra Costa County.<br />

(which dissipated energy from <strong>the</strong> 10% gradient) and <strong>the</strong><br />

banks were planted with willow (salix spp). Post-project<br />

appraisals <strong>in</strong> 1999 and 2004 (Purcell et al., 2002; Purcell,<br />

2004) showed that <strong>the</strong> biotic condition of <strong>the</strong> restored<br />

reach was measurably better than an unrestored control<br />

section upstream and that <strong>the</strong> biotic condition did not<br />

improve fur<strong>the</strong>r between 1999 (two years post-project)<br />

and 2004 (seven years post-project), <strong>in</strong>dicat<strong>in</strong>g <strong>the</strong> stream<br />

may have reached its biotic potential with<strong>in</strong> two years.<br />

Purcell et al. (2002) conducted an attitud<strong>in</strong>al survey of<br />

<strong>the</strong> residents with<strong>in</strong> one block of <strong>the</strong> daylighted section of<br />

Baxter Creek. Of <strong>the</strong> 45 responses received, most were<br />

positive overall about <strong>the</strong> restoration, but many expressed<br />

concerns that <strong>the</strong> willow trees, some of which had grown<br />

to over 6 m <strong>in</strong> height, blocked <strong>the</strong> view across <strong>the</strong> park<br />

and potentially provided hid<strong>in</strong>g places for burglars. In a<br />

repeat survey of <strong>the</strong> neighborhood <strong>in</strong> 2004 (n = 45),<br />

Purcell found that about half of those who had moved to<br />

<strong>the</strong> neighborhood after <strong>the</strong> completion of <strong>the</strong> restoration<br />

did not realize <strong>the</strong> creek had formerly been <strong>in</strong> an underground<br />

culvert. Nearly all respondents reported <strong>the</strong>y<br />

enjoyed liv<strong>in</strong>g near <strong>the</strong> creek, many cit<strong>in</strong>g <strong>the</strong> sounds of<br />

<strong>the</strong> water, aes<strong>the</strong>tics, or accessibility for children or dogs.<br />

69% perceived an improvement s<strong>in</strong>ce <strong>the</strong> restoration was<br />

completed; 31% said conditions had worsened. Overall,<br />

<strong>the</strong> project was successful <strong>in</strong> creat<strong>in</strong>g a vibrant stream<br />

corridor where formerly <strong>the</strong>re had been only a relatively<br />

sterile strip of lawn. The success of <strong>the</strong> project led to <strong>the</strong><br />

formation of <strong>the</strong> ‘Friends of Baxter Creek’, a group which<br />

subsequently supported two o<strong>the</strong>r restoration projects <strong>in</strong><br />

downstream reaches of Baxter Creek (Lisa Owens-Viani,<br />

personal communication, 2006).<br />

Figure 4.10 Baxter Creek <strong>in</strong> Po<strong>in</strong>sett Park, El Cerrito, California.<br />

Photo by Alison Purcell, April 2007, about 10 years after<br />

construction. Note height of willows, some exceed<strong>in</strong>g 6 m.


56 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

There was some negative reaction to ‘overgrown’<br />

vegetation (Purcell, 2004), which is not unusual <strong>in</strong> urban<br />

stream restorations <strong>in</strong> nor<strong>the</strong>rn California. Blackberry<br />

Creek <strong>in</strong> <strong>the</strong> Thousand Oaks School <strong>in</strong> Berkeley was<br />

removed from an underground culvert and replaced with<br />

an open channel <strong>in</strong> 1995. Some residents had negative<br />

reactions to <strong>the</strong> density of willow growth, which precluded<br />

access or see<strong>in</strong>g <strong>in</strong>to <strong>the</strong> stream channel. In plann<strong>in</strong>g<br />

<strong>the</strong> 1990 restoration of Cortland Creek <strong>in</strong> Oakland,<br />

<strong>the</strong> plans by creek activists to extensively plant willows<br />

(for habitat) met with resistance from local residents, who<br />

did not want to create a place where crim<strong>in</strong>als could hide<br />

(Walter Hood, personal communication, 1995). Though<br />

not as well documented as <strong>the</strong> Baxter Creek case, many<br />

o<strong>the</strong>r urban stream restoration projects have been marked<br />

by similar confl icts between plant<strong>in</strong>g willows to enhance<br />

habitat and <strong>the</strong> desire by residents and police to see <strong>in</strong>to<br />

<strong>the</strong> channel to discourage crim<strong>in</strong>als. This has been true<br />

especially <strong>in</strong> low-<strong>in</strong>come neighborhoods, where concern<br />

about crime may be greater.<br />

4.6.2 Marsh Creek, Brentwood, California<br />

Marsh Creek dra<strong>in</strong>s 332 km2 , with its upper bas<strong>in</strong> mostly<br />

woodland, rangeland and farmland, and its lower 15 km<br />

travers<strong>in</strong>g a broad alluvial fan, which now supports <strong>the</strong><br />

urban areas of Brentwood and Oakley, about 60 km nor<strong>the</strong>ast<br />

of San Francisco (Figure 4.9). As typical of Mediterranean-climate<br />

streams, runoff from <strong>the</strong> catchment was<br />

naturally <strong>in</strong>termittent <strong>in</strong> all but wet years. There is little<br />

record of <strong>the</strong> historical channel conditions <strong>in</strong> Marsh Creek<br />

<strong>in</strong> Brentwood, but historical maps from <strong>the</strong> late 1800s and<br />

early 1900s show multiple, s<strong>in</strong>uous channels and active<br />

channel migration (Rob<strong>in</strong>s and Ca<strong>in</strong>, 2002) (Figure 4.11).<br />

As agriculture expanded onto <strong>the</strong> fertile soils <strong>in</strong> <strong>the</strong> early–<br />

mid 20th century, and <strong>in</strong> response to fl ood<strong>in</strong>g of downtown<br />

Brentwood <strong>in</strong> <strong>the</strong> 1950s, <strong>the</strong> channel of Marsh Creek was<br />

straightened, <strong>the</strong> riparian corridor largely cleared and a<br />

fl ood-control reservoir constructed about 3 km upstream<br />

of Brentwood (Figure 4.12).<br />

Brentwood has grown rapidly, <strong>in</strong>creas<strong>in</strong>g <strong>in</strong> population<br />

from 7500 <strong>in</strong> 1990, to 23 000 <strong>in</strong> 2000, to 33 000 <strong>in</strong> 2003<br />

(Ca<strong>in</strong> et al., 2003). Many of <strong>the</strong>se residents commute (oneway<br />

travel times of over an hour) to jobs <strong>in</strong> <strong>the</strong> San Francisco<br />

Bay Region. As Brentwood has grown, <strong>in</strong>terest has<br />

grown <strong>in</strong> enhanc<strong>in</strong>g <strong>the</strong> creek corridor for human uses,<br />

remov<strong>in</strong>g barriers to salmonid migration and improv<strong>in</strong>g<br />

stormwater detention. A watershed study (Ca<strong>in</strong> et al., 2003)<br />

documented historical changes <strong>in</strong> physical and biological<br />

conditions, identify<strong>in</strong>g signifi cant effects of straighten<strong>in</strong>g<br />

on channel form and <strong>in</strong>stream habitat, effects of <strong>the</strong> altered<br />

fl ow regime on habitat, effects of former mercury m<strong>in</strong><strong>in</strong>g<br />

upstream and urban/agricultural runoff on water quality<br />

and <strong>the</strong> loss of native plant and animal species.<br />

To better understand <strong>the</strong> perceptions and preferences of<br />

local residents, Yang (2004) surveyed 1800 residents<br />

liv<strong>in</strong>g with<strong>in</strong> 400 m of Marsh Creek <strong>in</strong> Brentwood to<br />

assess <strong>the</strong>ir perceptions (and ideal images) of <strong>the</strong> creek.<br />

The residents consistently presented an ideal image of <strong>the</strong><br />

creek, identify<strong>in</strong>g luxuriant woods, year-round runn<strong>in</strong>g<br />

water, bountiful wildlife and easy access as features of <strong>the</strong><br />

‘natural’ or ‘orig<strong>in</strong>al’ Marsh Creek. However, this idyllic<br />

image of <strong>the</strong> creek is largely <strong>in</strong>consistent with <strong>the</strong> character<br />

of <strong>the</strong> creek as documented by historical evidence, with<br />

its Mediterranean-climate runoff regime. Similarly, many<br />

residents delighted <strong>in</strong> contact with wildlife, but did not<br />

realize that <strong>the</strong> most contacted species, i.e. crayfi sh, bullfrogs,<br />

bluegill and largemouth bass, are not native, but<br />

exotic generalists. Likewise, with vegetation, most subjects<br />

did not dist<strong>in</strong>guish native from <strong>in</strong>troduced plant<br />

species and even those who could tended to prefer vegetation<br />

that ‘looks natural without be<strong>in</strong>g overgrown’ regardless<br />

of orig<strong>in</strong> (Yang, 2004).<br />

The problems identifi ed by <strong>the</strong> residents contrasted<br />

sharply with those presented by <strong>the</strong> professionals <strong>in</strong> <strong>the</strong><br />

watershed report. By far <strong>the</strong> lead<strong>in</strong>g concern of surveyed<br />

residents was garbage and dump<strong>in</strong>g <strong>in</strong> <strong>the</strong> creek. Many<br />

residents considered <strong>the</strong> summer water levels too low,<br />

evidently without understand<strong>in</strong>g <strong>the</strong> highly seasonal nature<br />

of fl ow <strong>in</strong> Mediterranean-climate streams. Residents also<br />

regarded ‘mosquitoes/pests’ as more serious than ‘monotonous<br />

channel form’ and ‘poor habitat value,’ both major<br />

issues identifi ed <strong>in</strong> <strong>the</strong> watershed report (Ca<strong>in</strong> et al.,<br />

2003). Only ‘not enough shade’ was identifi ed as a concern<br />

both by <strong>the</strong> surveyed residents and <strong>in</strong> <strong>the</strong> watershed report<br />

(Yang, 2004).<br />

The substantial differences <strong>in</strong> perception and landscape<br />

preference between restoration scientists and local residents<br />

will be a source of uncerta<strong>in</strong>ty <strong>in</strong> sett<strong>in</strong>g restoration<br />

priorities and garner<strong>in</strong>g public support for restoration<br />

projects. As fund<strong>in</strong>g becomes available to plan restoration<br />

projects <strong>in</strong> Brentwood, <strong>the</strong>se gaps will need to be addressed<br />

<strong>in</strong> a participatory context so that confl icts <strong>in</strong> restoration<br />

goals can be reduced.<br />

4.7 CONCLUSIONS<br />

Uncerta<strong>in</strong>ties on <strong>the</strong> social and cultural fronts of stream<br />

restoration can be viewed as signify<strong>in</strong>g forward progress<br />

<strong>in</strong> <strong>the</strong> fi eld, ra<strong>the</strong>r than simply fur<strong>the</strong>r impediments<br />

<strong>in</strong> implement<strong>in</strong>g projects. Twenty years ago, when <strong>the</strong><br />

notions of creek restoration fi rst became widespread <strong>in</strong> <strong>the</strong><br />

United States, many eng<strong>in</strong>eers regarded ecological concerns<br />

as obstacles to be overcome <strong>in</strong> <strong>the</strong> s<strong>in</strong>gle-m<strong>in</strong>ded


a 1 0.5 0<br />

1 km b<br />

Plann<strong>in</strong>g <strong>River</strong> <strong>Restoration</strong> Projects: Social and Cultural Dimensions 57<br />

Figure 4.11 Topographic map details of marsh Creek <strong>in</strong> Brentwood: (a) 1914 and (b) 1978. (Source: US Geological Survey topographic<br />

maps.)<br />

Figure 4.12 (See also colour plate section) View of Marsh Creek<br />

channel <strong>in</strong> Brentwood (Photo by Kondolf, September 1991.)<br />

pursuit of dik<strong>in</strong>g, channeliz<strong>in</strong>g, straighten<strong>in</strong>g and culvert<strong>in</strong>g<br />

streams. It was only when eng<strong>in</strong>eers started to confront<br />

o<strong>the</strong>r viewpo<strong>in</strong>ts that ‘uncerta<strong>in</strong>ties’ were <strong>in</strong>troduced<br />

<strong>in</strong> <strong>the</strong>ir modus operandi. While confl icts between eng<strong>in</strong>eers<br />

and ecologists persist <strong>in</strong> restoration projects, by and<br />

large <strong>the</strong> eng<strong>in</strong>eer<strong>in</strong>g profession has embraced <strong>the</strong> need to<br />

work effectively with geomorphologists and biologists<br />

to achieve effective ecosystem restoration. Now we see<br />

<strong>in</strong>creas<strong>in</strong>gly that <strong>the</strong> ecological eng<strong>in</strong>eer<strong>in</strong>g approach is<br />

perturbed by <strong>the</strong> uncerta<strong>in</strong>ties <strong>in</strong>troduced by social and<br />

cultural concerns. In o<strong>the</strong>r words, <strong>the</strong> current phenomenon<br />

of restoration professionals experienc<strong>in</strong>g uncerta<strong>in</strong>ties on<br />

all fronts may be simply an <strong>in</strong>dicator of a rapidly broaden<strong>in</strong>g<br />

viewpo<strong>in</strong>t and recognition of problems without commensurate<br />

solutions.


58 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Although <strong>the</strong> possible range of goals, values, perception,<br />

aes<strong>the</strong>tic taste, use and mean<strong>in</strong>gs of a population for<br />

a stream can be overwhelm<strong>in</strong>g, and confl icts sometimes<br />

unavoidable, <strong>the</strong> goal of susta<strong>in</strong>able stream restoration<br />

<strong>in</strong>creas<strong>in</strong>gly requires <strong>in</strong>tegration of diverse po<strong>in</strong>ts of view.<br />

Institutionally, citizen groups have now become an <strong>in</strong>tegral<br />

part of many restoration efforts. Professionals have<br />

learnt that collaborative plann<strong>in</strong>g processes may not br<strong>in</strong>g<br />

about a fast solution, but it may provide substantial long<br />

term benefi ts <strong>in</strong> terms of political and social capital, and<br />

may yield more susta<strong>in</strong>able restoration projects.<br />

To be successful on a susta<strong>in</strong>able basis, stream restorations<br />

must be both technically sound and enjoy strong<br />

public support. Although decisions <strong>in</strong> stream restoration<br />

are essentially value driven, sound science is fundamental<br />

to constra<strong>in</strong> <strong>the</strong> range of possible solutions and evaluate<br />

possible alternatives. Without it, a Jamison Creek situation<br />

can result, <strong>in</strong> which <strong>the</strong> responsible agency selects a<br />

scientifi cally unsound option and <strong>the</strong> project fails. On <strong>the</strong><br />

o<strong>the</strong>r hand, a technically sound restoration plan is unlikely<br />

to be funded and implemented without strong public<br />

support, and unlikely to be susta<strong>in</strong>able if built without<br />

local buy-<strong>in</strong>.<br />

Where <strong>the</strong>re are signifi cant uncerta<strong>in</strong>ties on social and<br />

cultural aspects, <strong>the</strong>se should probably be settled before<br />

proceed<strong>in</strong>g to settle technical uncerta<strong>in</strong>ties. For example,<br />

until <strong>the</strong> values of large woody debris for fi sh or boaters<br />

are established, <strong>the</strong>re may be little po<strong>in</strong>t <strong>in</strong> quantify<strong>in</strong>g<br />

its catchment production and morphological qualities. In<br />

cities, we fi nd that <strong>the</strong> recreational potential of spontaneous<br />

uses is often conspicuous <strong>in</strong> its absence from <strong>the</strong><br />

agenda of stream restoration. Once <strong>the</strong>ir importance is<br />

recognized and spontaneous uses and <strong>the</strong>ir implied societal<br />

values are added to <strong>the</strong> restoration agenda, more<br />

precise research may be needed to assess <strong>the</strong>m.<br />

Cultural preferences (commonly unacknowledged)<br />

largely shape restoration goals. Build<strong>in</strong>g a culturally preferred<br />

form (such as a stable, meander<strong>in</strong>g channel) is<br />

perfectly reasonable as a restoration goal, but we suspect<br />

that <strong>the</strong> fi eld would benefi t from an explicit recognition of<br />

this as motivation, ra<strong>the</strong>r than cloak<strong>in</strong>g such projects <strong>in</strong><br />

seem<strong>in</strong>gly scientifi c details of channel morphology and<br />

(commonly vague) references to improved fi sh habitat. To<br />

<strong>the</strong> degree that cultural preferences rema<strong>in</strong> unacknowledged,<br />

<strong>the</strong>y <strong>in</strong>troduce greater uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> trajectory<br />

of restoration projects. Cultural preference for tidy landscapes<br />

over messy landscapes (Piégay et al., 2005) should<br />

be acknowledged, so that ‘overgrown’ riparian zones can<br />

ei<strong>the</strong>r be ‘framed’ (Nassauer, 1995) or simply avoided <strong>in</strong><br />

urban areas. For ecological design to be truly successful<br />

and widely accepted, designers will need to fi nd ways to<br />

make stream restoration compell<strong>in</strong>g as designs (Moz<strong>in</strong>go,<br />

1997). The concept of ‘eco-revelatory design’ suggests<br />

that by accept<strong>in</strong>g humans <strong>in</strong>to <strong>the</strong> restored ecosystem and<br />

design<strong>in</strong>g <strong>the</strong> project to reveal ecological processes, we<br />

may achieve ecosystem restoration (to <strong>the</strong> extent possible<br />

<strong>in</strong> urban areas) while still ga<strong>in</strong><strong>in</strong>g public acceptance<br />

(Galatowitsch, 1998).<br />

4.8 ACKNOWLEDGEMENTS<br />

The research on which this chapter is based was partially<br />

supported by a grant from <strong>the</strong> University of California,<br />

Berkeley, Department of Landscape Architecture Beatrix<br />

Farrand Fund. Shannah Anderson contributed substantially<br />

with support<strong>in</strong>g research, fi gure and manuscript<br />

preparation, and review comments. Louise Moz<strong>in</strong>go contributed<br />

valuable ideas and references. The chapter was<br />

improved through comments by anonymous reviewers<br />

and <strong>the</strong> volume’s editors, Dave Sear and Steve Darby.<br />

REFERENCES<br />

Appleton J. 1975. The Experience of Landscape. John Wiley &<br />

Sons Ltd: Chichester.<br />

Bauer DM, Cyr NE, Swallow SK. 2002. Public preferences for<br />

compensatory mitigation of salt marsh losses: a cont<strong>in</strong>gent<br />

choice of alternatives. Conservation Biology 18: 401–411.<br />

Bernhardt ES et al. 2005. Syn<strong>the</strong>siz<strong>in</strong>g US river restoration<br />

efforts. Science 308: 636–637.<br />

Brookes A, Shields FD. 1996. <strong>River</strong> Channel <strong>Restoration</strong>:<br />

Guid<strong>in</strong>g Pr<strong>in</strong>ciples For Susta<strong>in</strong>able Projects. John Wiley &<br />

Sons Ltd: Chichester.<br />

Ca<strong>in</strong> JR, Rob<strong>in</strong>s JD, Beamish SS. 2003. The Past and Present<br />

Condition of <strong>the</strong> Marsh Creek Watershed. Natural Heritage<br />

Institute: San Francisco, California.<br />

Carson R. 1962. Silent Spr<strong>in</strong>g. Houghton Miffl <strong>in</strong> Company: New<br />

York, New York.<br />

CCHR (Citizens Commission on Human Rights). 2004. The<br />

real crisis on mental health today. Available at Community<br />

Ru<strong>in</strong>: Psychiatry’s Coercive Care, CCHR: Los Angeles,<br />

California.<br />

Chawla L. 1988. Children’s concern for <strong>the</strong> natural environment.<br />

Children’s Environment Quarterly 5 (3):13–20.<br />

Cobb E. 1977. The Ecology of Imag<strong>in</strong>ation <strong>in</strong> Childhood. Columbia<br />

University Press: New York, New York.<br />

Connick S, Innes J. 2003. Outcomes of collaborative water policy<br />

mak<strong>in</strong>g: apply<strong>in</strong>g complexity th<strong>in</strong>k<strong>in</strong>g to evaluation. Journal<br />

of Environmental Plann<strong>in</strong>g and Management 46: 177–197.<br />

Cooper-Marcus C. 1992. Environmental memories. In: Low SM,<br />

Altman I (Eds), Place Attachment, Human Behavior and Environment:<br />

Advances <strong>in</strong> Theory and Research, Vol. 12. Plenum<br />

Press: New York, New York; 87–112.<br />

Daniel TC, V<strong>in</strong><strong>in</strong>g J. 1983. Methodological issues <strong>in</strong> <strong>the</strong> assessment<br />

of landscape quality. In: Altman I, Wohlwill JF (Eds),<br />

Behavior and Natural Environment, Plenum Press: New York,<br />

New York; 39–84.


Dohi M. 1999. The community design process at Kamagasaki,<br />

Osaka, Japan. In: Hester RT, Kwesk<strong>in</strong> C (Eds), Democratic<br />

Design <strong>in</strong> <strong>the</strong> Pacifi c Rim, Ridge Times Press: Mendoc<strong>in</strong>o,<br />

California; 228–241.<br />

Eden SS, Tunstall S, Tapsell S. 2000. Translat<strong>in</strong>g nature: river<br />

restoration as nature-culture. Environment and Plann<strong>in</strong>g D:<br />

Society and Space 18: 257–273.<br />

Ehrlich PR. 1968. The Population Bomb. Buccaneer Books: New<br />

York, New York.<br />

Fagan K. 2005. Solano’s hidden homeless. San Francisco Chronicle.<br />

13 November 2005, pp. 1, A10–11.<br />

FISCRWG (Federal Interagency Stream Corridor <strong>Restoration</strong><br />

Work<strong>in</strong>g Group). 1998. Stream Corridor <strong>Restoration</strong> Handbook.<br />

Natural Resources Conservation Service: Wash<strong>in</strong>gton,<br />

DC.<br />

Flosi G et al. 1998. California Salmonid Stream Habitat <strong>Restoration</strong><br />

Manual (3rd edition). California Department of Fish and<br />

Game: Sacramento, California.<br />

Fornes J. 1994. Plan for <strong>the</strong> Metropolitan Park of Havana.<br />

Parque Metropolitano de la Habana: Havana, Cuba.<br />

Galatowitsch SM. 1998. Ecological design for environmental<br />

problem solv<strong>in</strong>g. Landscape Journal 17: 99–107.<br />

Garcia and Associates. 2005. Rock Creek-Cresta recreational<br />

streamfl ow monitor<strong>in</strong>g year one report: 2002 macro<strong>in</strong>vertebrate<br />

drift sampl<strong>in</strong>g North Fork Fea<strong>the</strong>r <strong>River</strong>, Plumas County<br />

CA. Prepared for Pacifi c Gas and Electric Company (FERC<br />

No. 1962).<br />

Hamilton J. 1993. Streams of hope. Sierra September/October<br />

1993: 98–122.<br />

Hart R. 1979. Children’s Experience of Place. Irv<strong>in</strong>gton Publishers:<br />

New York, New York.<br />

Harvey MR. 1989. Children’s experiences with vegetation.<br />

Children’s Environment Quarterly 6 (1): 36–43.<br />

Hester RT et al. 1988. ‘We’d like to tell you . . .’: Children’s<br />

views of life <strong>in</strong> Westport, California. Small Town 18 (4):<br />

19–24.<br />

Hogarth W. 1753. The Analysis of Beauty. Repr<strong>in</strong>ted <strong>in</strong> 1997 by<br />

Yale University Press: New Haven.<br />

Hough M. 1990. Out of Place: Restor<strong>in</strong>g Identity to <strong>the</strong> Regional<br />

Landscape. Yale University Press, New Haven.<br />

Hull RB, Robertson DP. 2000. The language of nature matters:<br />

we need a more public ecology. In: Gobster PH, Hull RB<br />

(Eds), Restor<strong>in</strong>g Nature, Island Press: Wash<strong>in</strong>gton, DC;<br />

7–118.<br />

Jones DR, Battaglia AM. 1989. Ma<strong>in</strong> Street <strong>River</strong>s: Mak<strong>in</strong>g Connections<br />

Between <strong>River</strong>s and Towns. Pennsylvania State University:<br />

State College, Pennsylvania.<br />

Junk WJ, Bayley PB, Sparks RE. 1989. The fl ood pulse concept<br />

<strong>in</strong> river-fl oodpla<strong>in</strong> system. Canadian Journal of Fisheries and<br />

Aquatic Sciences. 106: 110–127.<br />

Kaplan R. 1977. Preference and everyday nature: method and<br />

application. In: Stokols D (Ed), Perspectives on Environment<br />

and Behavior: Theory, Research, and Applications, Plenum<br />

Press: New York, New York; 235–250.<br />

Kaplan R, Kaplan S. 1989. The Experience of Nature: A Psychological<br />

Perspective. Cambridge University Press: New York,<br />

New York.<br />

Plann<strong>in</strong>g <strong>River</strong> <strong>Restoration</strong> Projects: Social and Cultural Dimensions 59<br />

Kondolf GM, Smeltzer MW, Railsback S. 2001. Design and<br />

performance of a channel reconstruction project <strong>in</strong> a coastal<br />

California gravel-bed stream. Environmental Management 28<br />

(6): 761–776.<br />

Lewis SA. 1995. Design and plann<strong>in</strong>g implications of uses, perceptions<br />

and attitudes of San Leandro Creek. Master’s <strong>the</strong>sis.<br />

Univeristy of California, Berkeley.<br />

Marchetti MP, Moyle PB. 2001. Effects of fl ow regime on fi sh<br />

assemblages <strong>in</strong> a regulated California stream. Ecological<br />

Applications 11 (2): 530–539.<br />

Marzolf GR, Valdez RA, Schmidt JC, Webb RH. 1998. Perspectives<br />

on river restoration <strong>in</strong> <strong>the</strong> Grand Canyon. Bullet<strong>in</strong> of <strong>the</strong><br />

Ecological Society of America 79 (4): 250–254.<br />

Moore RC. 1986. Childhood’s Doma<strong>in</strong>. Croom Helm: London.<br />

Moz<strong>in</strong>go LA. 1997. The aes<strong>the</strong>tics of ecological design: see<strong>in</strong>g<br />

science as culture. Landcape Journal 16: 46–59.<br />

Moz<strong>in</strong>go LA. 2005. Community participation and creek restoration<br />

<strong>in</strong> <strong>the</strong> East Bay of San Francisco, California. In: Hou J,<br />

Francis M, Brightbill N (Eds), (Re)construct<strong>in</strong>g Communities,<br />

<strong>the</strong> 5th Pacifi c Rim Conference on Participatory Community<br />

Design, Center for Design Research: University of California,<br />

Davis; 249–251.<br />

Nassauer JI. 1995. Messy ecosystems, orderly frames. Landscape<br />

Journal 14: 161–170.<br />

Nicholson S. 1971. The <strong>the</strong>ory of loose parts. Landscape Architecture<br />

62 (1): 30–34.<br />

NRC (National Research Council). 1992. <strong>Restoration</strong> of Aquatic<br />

Ecosystems. National Academy Press: Wash<strong>in</strong>gton, DC.<br />

Orr D. 1992. Ecological Literacy: Education and <strong>the</strong> Transition<br />

to a Postmodern World. State University of New York Press:<br />

Albany, New York.<br />

Otto B, McCormick K, Leccese M. 2004. Ecological riverfront<br />

design: restor<strong>in</strong>g rivers, connect<strong>in</strong>g communities. American<br />

Plann<strong>in</strong>g Association Plann<strong>in</strong>g Advisory Service Report<br />

Number 518–519, American Plann<strong>in</strong>g Association: Chicago,<br />

Ill<strong>in</strong>ois.<br />

Owens PE. 1988. Natural landscapes, ga<strong>the</strong>r<strong>in</strong>g places, and prospect<br />

refuges: characteristics of outdoor places valued by teens.<br />

Children’s Environments Quarterly 5 (2): 17–24.<br />

Piégay H, Mutz M, Gregory KJ. 2005. Public perception as a<br />

barrier to <strong>in</strong>troduc<strong>in</strong>g wood <strong>in</strong> rivers for restoration purposes.<br />

Environmental Management 36 (5): 665–674.<br />

Poff NL et al. 1997. The natural fl ow regime. BioScience 47:<br />

769–784.<br />

Postel S, Richter B. 2004. <strong>River</strong>s for Life. Island Press: Covelo,<br />

California.<br />

Purcell AH, Friedrich C, Resh VH. 2002. An assessment of a<br />

small urban stream restoration project <strong>in</strong> nor<strong>the</strong>rn California.<br />

<strong>Restoration</strong> Ecology 10: 685–694.<br />

Purcell AH. 2004. A long term post-project evaluation of an<br />

urban stream restoration project (Baxter Creek, el Cerrito, California).<br />

Term project <strong>in</strong> <strong>River</strong> <strong>Restoration</strong>, LA227, University of<br />

California, Berkeley. Available through <strong>the</strong> University of California<br />

library website (http://repositories.cdlib.org/wrca/<br />

restoration/)<br />

Reiser DW, Bjornn TC. 1979. Habitat requirements of anadromous<br />

salmonids. USDA General Technical Report PNW-96.


60 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Riley AL. 1998. Restor<strong>in</strong>g Streams <strong>in</strong> Cities: A Guide for Planners,<br />

Policymakers, and Citizens. Island Press: Wash<strong>in</strong>gton,<br />

DC.<br />

Rob<strong>in</strong>s JD, Ca<strong>in</strong> JR. 2002. The Past and Present Condition of <strong>the</strong><br />

Marsh Creek Watershed. Report by <strong>the</strong> Natural Heritage Institute<br />

and <strong>the</strong> Delta Science Center to <strong>the</strong> California Coastal<br />

Conservancy: Oakland, California.<br />

Rood SB, Mahoney JM. 2000. Revised <strong>in</strong>stream fl ow regulation<br />

enables cottonwood recruitment along <strong>the</strong> St Mary <strong>River</strong>,<br />

Alberta, Canada. <strong>River</strong>s 7: 109–125.<br />

Rosgen DL. 1994. A classifi cation of natural rivers. Catena 22:<br />

169–199.<br />

Ryan RL. 2000. A people-centered approach to restoration projects:<br />

<strong>in</strong>sights from understand<strong>in</strong>g attachment to urban natural<br />

areas. In: Gobster PH, Hull RB (Eds), Restor<strong>in</strong>g Nature, Island<br />

Press: Wash<strong>in</strong>gton, DC; 209–228.<br />

Ryan RL, Erickson DL, DeYoung R. 2002. Farmers’ motivations<br />

for adopt<strong>in</strong>g conservation practices along riparian zones <strong>in</strong> a<br />

Midwestern agricultural watershed. Journal of Environmental<br />

Plann<strong>in</strong>g and Management 46 (1): 19–37.<br />

Schmidt JC et al. 1998. Science and values <strong>in</strong> river restoration <strong>in</strong><br />

<strong>the</strong> Grand Canyon. BioScience 48 (9): 735–747.<br />

Smith SM, Prestegaard KL. 2005. Hydraulic performance of a<br />

morphology-based stream channel design. Water Resources<br />

Research 41 (W11413): 1–17.<br />

Tunstall SM, Penn<strong>in</strong>g-Rowsell EC, Tapsell SM, Eden SE. 2000.<br />

<strong>River</strong> restoration: public attitudes and expectations. Journal of<br />

<strong>the</strong> Institution of Water and Environmental Management 14<br />

(5): 363–370.<br />

Tunstall SM, Tapsell SM, House M. 2004. Children’s perceptions<br />

of river landscapes and play: what children’s photographs<br />

reveal. Landscape Research 29 (2): 181–204.<br />

Ulrich RS. 1983. Aes<strong>the</strong>tic and effective response to<br />

natural environment. In: Altman I, Wohlwill J (Eds), Behavior<br />

and <strong>the</strong> Natural Environment, Plenum: New York, New<br />

York.<br />

Ward JV, Stanford JA. 1995. Ecological connectivity <strong>in</strong> alluvial<br />

river ecosystems and its disruption by fl ow regulation. Regulated<br />

rivers: Research and Management 11: 105–119.<br />

Wolman MG. 1971. The nation’s rivers. Science 174: 905–918.<br />

Yang C-N. 2004. Invit<strong>in</strong>g Spontaneous Use <strong>in</strong>to Urban Streams.<br />

Doctoral dissertation. University of California, Berkeley.<br />

Available onl<strong>in</strong>e at http://www.lib.berkeley.edu/WRCA/<br />

restoration/<strong>the</strong>ses.html.


<strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat Edited by Stephen Darby and David Sear<br />

© 2008 John Wiley & Sons, Ltd<br />

5<br />

Conceptual and Ma<strong>the</strong>matical Modell<strong>in</strong>g <strong>in</strong><br />

<strong>River</strong> <strong>Restoration</strong>: Do We Have<br />

Unreasonable Confi dence?<br />

Michael Stewardson 1 and Ian Ru<strong>the</strong>rfurd 2<br />

1 Department of Civil and Environmental Eng<strong>in</strong>eer<strong>in</strong>g and eWater CRC, The University of Melbourne, Australia<br />

2 Geography Program, School of Resource Management, The University of Melbourne, Australia<br />

5.1 INTRODUCTION<br />

5.1.1 Geomorphic Modell<strong>in</strong>g <strong>in</strong><br />

<strong>Restoration</strong> Plann<strong>in</strong>g<br />

From <strong>the</strong> 1960s to <strong>the</strong> 1980s, applied fl uvial geomorphology<br />

described <strong>the</strong> degradation of streams <strong>in</strong> response to<br />

human disturbance. This endeavor was part of a larger<br />

quest to expla<strong>in</strong> <strong>the</strong> controls on stream form. From <strong>the</strong><br />

1990s, <strong>the</strong> discipl<strong>in</strong>e has found new vigor as skills have<br />

been turned to <strong>the</strong> restoration of those degraded streams.<br />

However, this change does not simply represent a new job<br />

opportunity; it represents a fundamental test of our knowledge<br />

of processes controll<strong>in</strong>g stream form.<br />

It is relatively easy to describe <strong>the</strong> degradation of fl uvial<br />

systems, and much of this work has identifi ed associations<br />

with human disturbances ra<strong>the</strong>r than strict causations.<br />

Recall <strong>the</strong> protracted debates about whe<strong>the</strong>r arroyo <strong>in</strong>cision<br />

was caused by clear<strong>in</strong>g, channelisation or climate<br />

change (Cooke and Reeves, 1976). It is much more challeng<strong>in</strong>g,<br />

fi rstly to recommend priorities for rehabilitat<strong>in</strong>g<br />

streams and, secondly, to implement <strong>the</strong>se changes. Geomorphologists<br />

have moved from be<strong>in</strong>g observers of human<br />

impact, to advisors and managers, actively <strong>in</strong>terven<strong>in</strong>g <strong>in</strong><br />

stream channels for environmental outcomes. Most of this<br />

<strong>in</strong>tervention has been <strong>in</strong> <strong>the</strong> areas of: channelised streams<br />

(Larson and Goldsmith, 1997), sediment slugs (Ru<strong>the</strong>rfurd,<br />

2001) and mitigat<strong>in</strong>g <strong>the</strong> effects of dams and fl ow<br />

regulation. Numerous articles describe <strong>the</strong> contribution<br />

that geomorphologists can make to stream restoration<br />

endeavors and <strong>the</strong>re is no shortage of admonitions to<br />

‘<strong>in</strong>clude a geomorphologist on every restoration project’<br />

(Sear, 1994; Brookes, 1995; Brierley et al., 1996; Brookes<br />

and Sear, 1996) and on every stream eng<strong>in</strong>eer<strong>in</strong>g project<br />

(Gilvear, 1999).<br />

Fluvial geomorphologists produce conceptual and<br />

ma<strong>the</strong>matical models that are central to many stream restoration<br />

projects. In this chapter it is argued that managers,<br />

ecologists, and even geomorphologists <strong>the</strong>mselves, can<br />

have a false sense of confi dence <strong>in</strong> <strong>the</strong>se models. As communities<br />

around <strong>the</strong> world <strong>in</strong>vest <strong>in</strong> stream restoration<br />

projects, false confi dence can have a huge direct and<br />

opportunity cost. We believe it is better to be frank about<br />

model uncerta<strong>in</strong>ties from <strong>the</strong> outset to promote realistic<br />

expectations of project success and a balanced and welltargeted<br />

<strong>in</strong>vestment <strong>in</strong> <strong>in</strong>vestigations to reduce <strong>the</strong>se<br />

uncerta<strong>in</strong>ties. Fluvial geomorphologists contribute to<br />

stream restoration projects <strong>in</strong> three ma<strong>in</strong> ways:<br />

1. The most basic work of a fl uvial geomorphologist <strong>in</strong><br />

a restoration project is to describe how a stream has<br />

changed its form over time, and to identify <strong>the</strong> factors<br />

that are responsible for <strong>the</strong>se changes. Such reconstructions<br />

have now passed from be<strong>in</strong>g a curiosity-driven<br />

activity for scientists, to basic consult<strong>in</strong>g practice. The<br />

result of <strong>the</strong>se <strong>in</strong>vestigations is usually a conceptual<br />

model [or a perceptual model <strong>in</strong> <strong>the</strong> parlance of Beven<br />

(2001)] that may be based on space-for-time substitution<br />

or historical co<strong>in</strong>cidence between geomorphic and


62 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

o<strong>the</strong>r catchment changes. Perhaps <strong>the</strong> best established<br />

geomorphic conceptual models are <strong>the</strong> <strong>in</strong>cised stream<br />

models (Schumm et al., 1984; Simon, 1989) where<br />

<strong>the</strong>re may be some empirical guidance for <strong>the</strong> magnitude<br />

of change (e.g. a threshold width–depth ratio for<br />

<strong>in</strong>cised streams Schumm et al., 1984). The conceptual<br />

model serves to describe <strong>the</strong> ‘orig<strong>in</strong>al’ condition of a<br />

system, which often becomes <strong>the</strong> ‘target’ for <strong>the</strong> restoration<br />

project. Fur<strong>the</strong>r, <strong>the</strong> conceptual model is used to<br />

predict <strong>the</strong> trajectory of channel change if <strong>the</strong>re is no<br />

<strong>in</strong>tervention.<br />

2. Geomorphologists recommend specifi c actions (or<br />

<strong>in</strong>terventions) that will alter process and form, to move<br />

<strong>the</strong> stream toward <strong>the</strong> ‘target’ state, identifi ed from <strong>the</strong><br />

conceptual model. These actions usually <strong>in</strong>volve chang<strong>in</strong>g<br />

<strong>the</strong> fl ow or sediment regime, or modify<strong>in</strong>g <strong>the</strong><br />

boundary of <strong>the</strong> channel to cause a change <strong>in</strong> some<br />

geomorphic variable (e.g. width, scour frequency,<br />

erosion rate). Design of project specifi cs is often based<br />

on a ma<strong>the</strong>matical model of geomorphic response<br />

related to <strong>the</strong> specifi c action proposed.<br />

3. They also predict <strong>the</strong> ‘secondary’ (perhaps un<strong>in</strong>tended)<br />

consequences of stream restoration projects. This<br />

could also be described as <strong>the</strong> ‘susta<strong>in</strong>ability’ of <strong>the</strong><br />

<strong>in</strong>tervention.<br />

Thus, uncerta<strong>in</strong>ty emerges at three levels: <strong>the</strong> validity<br />

of <strong>the</strong> conceptual model; whe<strong>the</strong>r <strong>the</strong> proposed <strong>in</strong>tervention<br />

results <strong>in</strong> <strong>the</strong> planned geomorphic change; and, fi nally,<br />

whe<strong>the</strong>r <strong>the</strong> change is susta<strong>in</strong>able. Typical examples of<br />

<strong>in</strong>-stream geomorphic actions <strong>in</strong> restoration projects are:<br />

• Incised, channelised streams are <strong>the</strong> classical example of<br />

stream restoration. A geomorphologist reconstructs<br />

<strong>the</strong> orig<strong>in</strong>al dimensions of a channelised stream <strong>in</strong><br />

Denmark by superimpos<strong>in</strong>g old maps and photos, and by<br />

mak<strong>in</strong>g pa<strong>in</strong>stak<strong>in</strong>g fi eld observations (Neilsen, 1996).<br />

Work<strong>in</strong>g with an eng<strong>in</strong>eer, a ‘stable’ re-meandered<br />

stream path that l<strong>in</strong>ks <strong>the</strong> palaeochannels for a rehabilitated<br />

stream is <strong>the</strong>n designed. The dimensions of <strong>the</strong><br />

channel and <strong>the</strong> variations <strong>in</strong> depth (<strong>the</strong> pool-riffl e<br />

sequence) are designed to be scaled to catchment area<br />

(Newbury and Gaboury, 1993). F<strong>in</strong>ally, <strong>the</strong> restoration<br />

that will lead to <strong>in</strong>creased erosion downstream of <strong>the</strong><br />

reach is predicted.<br />

• Gully<strong>in</strong>g has dumped a large pulse of sand <strong>in</strong>to a stream.<br />

A geomorphologist applies <strong>the</strong> classical ‘wave’ model<br />

of sand slug migration (Gilbert, 1917; Neilsen, 1996)<br />

and, on <strong>the</strong> basis of historical movement of <strong>the</strong> sand<br />

front, concludes that it will take over 50 years for <strong>the</strong><br />

sand to move through <strong>the</strong> reach. Extract<strong>in</strong>g <strong>the</strong> sand<br />

at a defi ned rate will protect <strong>the</strong> downstream reaches<br />

and accelerate recovery (Ru<strong>the</strong>rfurd, 2001). Build<strong>in</strong>g<br />

artifi cial spur dikes will also create habitat pools<br />

(Kuhnle et al., 2002).<br />

5.1.2 Why do we Care about <strong>Uncerta<strong>in</strong>ty</strong>?<br />

Regan et al. (2002) defi ne epistemic uncerta<strong>in</strong>ty as uncerta<strong>in</strong>ty<br />

associated with knowledge of <strong>the</strong> state of a system;<br />

it can be classifi ed <strong>in</strong>to six ma<strong>in</strong> types <strong>in</strong>clud<strong>in</strong>g random<br />

measurement error, natural variation and model uncerta<strong>in</strong>ty<br />

(compare to <strong>the</strong> classifi cation presented <strong>in</strong> Chapter<br />

3). This chapter is specifi cally concerned with <strong>the</strong>se three<br />

sources of uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> context of conceptual and<br />

ma<strong>the</strong>matical geomorphic models used <strong>in</strong> restoration projects.<br />

Measurements of various sorts are used to establish<br />

<strong>the</strong> state of a system. Measurements may be used directly<br />

(e.g. <strong>the</strong> diameter of an <strong>in</strong>dividual gra<strong>in</strong> of sediment), but<br />

can also be transformed us<strong>in</strong>g some k<strong>in</strong>d of calibration<br />

(e.g. stream discharge, which is often estimated from an<br />

observation of stage and transformed to discharge us<strong>in</strong>g a<br />

rat<strong>in</strong>g curve). Measurement error is <strong>the</strong> result of errors <strong>in</strong><br />

<strong>the</strong> direct measurement and subsequent transformations.<br />

Natural variation presents a challenge for observ<strong>in</strong>g <strong>the</strong><br />

true state of a system. To address uncerta<strong>in</strong>ty, assumptions<br />

are often made about <strong>the</strong> statistical properties of environmental<br />

variations. Model uncerta<strong>in</strong>ty can arise both from<br />

<strong>the</strong> choice of variables and processes to be represented <strong>in</strong><br />

<strong>the</strong> model and from <strong>the</strong> method used to represent <strong>the</strong> relations<br />

between variables.<br />

<strong>Uncerta<strong>in</strong>ty</strong> is defi ned here as <strong>the</strong> range of possible<br />

values for a model variable (<strong>in</strong>put or response). One part<br />

of uncerta<strong>in</strong>ty is accuracy, which is <strong>the</strong> difference between<br />

a measurement and <strong>the</strong> ‘true value’. Uncerta<strong>in</strong>ties <strong>in</strong><br />

response variables are <strong>the</strong> consequence of <strong>the</strong> need to<br />

choose a conceptual and ma<strong>the</strong>matical representation of<br />

river processes, and uncerta<strong>in</strong>ties <strong>in</strong> <strong>in</strong>put parameters for<br />

<strong>the</strong> model estimated from fi eld measurements, previous<br />

studies or by calibration. Uncerta<strong>in</strong>ties can also exist as a<br />

consequence of unknown future environmental conditions<br />

<strong>in</strong> particular climatic conditions. It is possible that restoration<br />

works are destroyed by an extreme fl ood event not<br />

considered <strong>in</strong> <strong>the</strong> plann<strong>in</strong>g process. Design models often<br />

consider a range of expected conditions based on conditions<br />

experienced <strong>in</strong> <strong>the</strong> past, but uncerta<strong>in</strong>ties <strong>in</strong> <strong>the</strong><br />

actual conditions over <strong>the</strong> life <strong>the</strong> project create fur<strong>the</strong>r<br />

uncerta<strong>in</strong>ty <strong>in</strong> geomorphic design.<br />

Discussion of uncerta<strong>in</strong>ty, <strong>in</strong> <strong>the</strong> realm of management,<br />

can quickly degenerate <strong>in</strong>to an unfocused quest for greater<br />

accuracy, greater precision, more samples and more effort.<br />

It is usually scientists who write about certa<strong>in</strong>ty and <strong>the</strong>y<br />

may have a completely different perspective on <strong>the</strong> issue


Conceptual and Ma<strong>the</strong>matical Modell<strong>in</strong>g <strong>in</strong> <strong>River</strong> <strong>Restoration</strong>: Do We Have Unreasonable Confi dence? 63<br />

to managers. The notion of ‘certa<strong>in</strong>ty’ is not simply a<br />

statistical/scientifi c issue; it is at <strong>the</strong> heart of <strong>the</strong> social<br />

process of decision mak<strong>in</strong>g. The implications of estimat<strong>in</strong>g<br />

geomorphic uncerta<strong>in</strong>ty can be unexpected to those<br />

focused primarily on improved confi dence <strong>in</strong> model prediction.<br />

Whilst more certa<strong>in</strong> predictions should improve<br />

<strong>the</strong> chance of project success, consideration of uncerta<strong>in</strong>ties<br />

does not <strong>in</strong> itself improve model performance and can<br />

reduce public support for a project.<br />

It is necessary to be clear about why we care about<br />

geomorphic uncerta<strong>in</strong>ty <strong>in</strong> stream restoration. In reality,<br />

most scientifi c geomorphologists revel <strong>in</strong> <strong>the</strong> uncerta<strong>in</strong>ties<br />

of river process and form. Most of our endeavour is<br />

directed at situations where accepted models do not work,<br />

where we are uncerta<strong>in</strong>. If it was all certa<strong>in</strong>, we would do<br />

someth<strong>in</strong>g else. But <strong>in</strong> <strong>the</strong> realm of management, uncerta<strong>in</strong>ty<br />

is seldom welcome. Whilst <strong>the</strong> enthusiasm to rehabilitate<br />

streams will not disappear, <strong>the</strong> <strong>in</strong>itial fl ush of<br />

community and government support for <strong>the</strong>se endeavors<br />

could be lost if geomorphologists and o<strong>the</strong>r scientists<br />

foster unrealistic expectations of <strong>the</strong>se projects. As geomorphologists<br />

engage with eng<strong>in</strong>eers and managers, <strong>the</strong>y<br />

must decide how to deal with uncerta<strong>in</strong>ty (Volkman,<br />

1999), when admitt<strong>in</strong>g to too much uncerta<strong>in</strong>ty can weaken<br />

support for a project and possibly stall it. However, underplay<strong>in</strong>g<br />

uncerta<strong>in</strong>ties will underm<strong>in</strong>e confi dence <strong>in</strong> geomorphic<br />

advice when some projects <strong>in</strong>evitably ‘fail’.<br />

Ano<strong>the</strong>r reason to care about uncerta<strong>in</strong>ty is that it provides<br />

<strong>the</strong> justifi cation for improved geomorphic <strong>in</strong>vestigations<br />

prior to completion of a restoration plan. As will be<br />

demonstrated, <strong>in</strong> some cases it may be relatively cheap and<br />

easy to reduce uncerta<strong>in</strong>ty, but it can also be very expensive<br />

to do so. A strong conclusion of this chapter is that,<br />

given <strong>the</strong> large cost of stream restoration projects, sound<br />

geomorphic advice has often tended to be undervalued.<br />

Large projects are sometimes launched on <strong>the</strong> basis of<br />

fl imsy conceptual models, possibly because <strong>the</strong>ir uncerta<strong>in</strong>ty<br />

has not been properly considered.<br />

F<strong>in</strong>ally, adaptive management is often proposed as <strong>the</strong><br />

only reasonable way forward <strong>in</strong> <strong>the</strong> face of uncerta<strong>in</strong> restoration<br />

outcomes (see Chapter 14). The adaptive approach<br />

is to treat <strong>the</strong> restoration project as an experiment designed<br />

to <strong>in</strong>form our knowledge of how rivers respond to restoration.<br />

This knowledge is used to improve river restoration<br />

decisions for <strong>the</strong> particular river and presumably elsewhere.<br />

Despite <strong>the</strong> frequent calls for adaptive management<br />

of river restoration, actual examples of success are rare for<br />

a number of reasons (Walters, 1997; Ladson and Argent,<br />

2002). We argue that systematic consideration of uncerta<strong>in</strong>ties<br />

<strong>in</strong> geomorphic modell<strong>in</strong>g is essential to <strong>the</strong><br />

application of adaptive approaches <strong>in</strong> river restoration.<br />

Management experiments and associated monitor<strong>in</strong>g need<br />

be targeted to reduc<strong>in</strong>g <strong>the</strong>se uncerta<strong>in</strong>ties if <strong>the</strong>y are to<br />

feed back effectively <strong>in</strong>to future restoration practice.<br />

5.1.3 Introduction to Case Studies<br />

Geomorphologists normally acknowledge <strong>the</strong> uncerta<strong>in</strong>ties<br />

<strong>in</strong> <strong>the</strong>ir models but it is rare for <strong>the</strong>se uncerta<strong>in</strong>ties to<br />

be quantifi ed or systematically exam<strong>in</strong>ed. There could be<br />

a perception that <strong>the</strong>se uncerta<strong>in</strong>ties are relatively small<br />

and have limited signifi cance <strong>in</strong> restoration decisions.<br />

There may be limited experience amongst geomorphologists<br />

<strong>in</strong> <strong>the</strong> quantitative aspects of uncerta<strong>in</strong>ty analysis<br />

which could discourage attempts to handle <strong>the</strong>se explicitly.<br />

There may also be a concern that quantify<strong>in</strong>g uncerta<strong>in</strong>ties<br />

will underm<strong>in</strong>e support for a project. In any case,<br />

<strong>the</strong>re is currently very little published <strong>in</strong>formation on <strong>the</strong><br />

scale of uncerta<strong>in</strong>ties <strong>in</strong> geomorphic studies for river<br />

restoration. <strong>Uncerta<strong>in</strong>ty</strong> analyses <strong>in</strong> case study projects are<br />

needed to <strong>in</strong>form discussion of how best to handle <strong>the</strong>se<br />

uncerta<strong>in</strong>ties <strong>in</strong> <strong>the</strong> future.<br />

This chapter exam<strong>in</strong>es uncerta<strong>in</strong>ties <strong>in</strong> geomorphic<br />

modell<strong>in</strong>g for two river restoration projects. The two case<br />

studies <strong>in</strong>volve, respectively, conceptual and design models<br />

for restoration plann<strong>in</strong>g. The scale of uncerta<strong>in</strong>ties and<br />

also <strong>the</strong> potential benefi ts of a systematic analysis of<br />

uncerta<strong>in</strong>ties <strong>in</strong> <strong>the</strong>se projects are exam<strong>in</strong>ed. Based on<br />

<strong>the</strong>se case studies, it is suggested how uncerta<strong>in</strong>ties might<br />

best be handled <strong>in</strong> <strong>the</strong> future by geomorphologists develop<strong>in</strong>g<br />

conceptual and design models for river restoration<br />

plann<strong>in</strong>g.<br />

The fi rst case study concerns <strong>the</strong> development of a<br />

conceptual model for plann<strong>in</strong>g Australia’s largest stream<br />

restoration project, on <strong>the</strong> Snowy <strong>River</strong>. The <strong>in</strong>fl uence of<br />

<strong>the</strong> conceptual model on proposed plans is exam<strong>in</strong>ed and,<br />

<strong>in</strong> particular, <strong>the</strong> implication of subsequent changes <strong>in</strong> <strong>the</strong><br />

geomorphic model. The second case study concerns <strong>the</strong><br />

design of a fl ush<strong>in</strong>g fl ow <strong>in</strong> <strong>the</strong> Goulburn <strong>River</strong>, Victoria,<br />

us<strong>in</strong>g a one-dimensional hydraulic modell<strong>in</strong>g approach. In<br />

this design problem, multiple sources of uncerta<strong>in</strong>ties are<br />

quantifi ed and <strong>the</strong> key uncerta<strong>in</strong>ties <strong>in</strong> design<strong>in</strong>g <strong>the</strong> fl ush<strong>in</strong>g<br />

fl ow identifi ed. A two-dimensional numerical model<br />

may have had some advantages over a one-dimensional<br />

model approach <strong>in</strong> this problem. However, a onedimensional<br />

model (similar to HEC RAS) is used because<br />

it is currently <strong>the</strong> standard approach used throughout <strong>the</strong><br />

<strong>in</strong>dustry. Thus, we are concerned with <strong>the</strong> uncerta<strong>in</strong>ty associated<br />

with current practice <strong>in</strong> stream restoration projects<br />

ra<strong>the</strong>r than with <strong>the</strong> level of certa<strong>in</strong>ty that is <strong>the</strong>oretically<br />

possible with detailed scientifi c study. It is often<br />

assumed that uncerta<strong>in</strong>ties associated with modell<strong>in</strong>g <strong>the</strong><br />

physical response of channels are small (c.f. biological<br />

responses, see Chapter 8). The case studies demonstrate


64 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

that uncerta<strong>in</strong>ties associated with <strong>the</strong> ‘accepted’ approaches<br />

to geomorphic modell<strong>in</strong>g can be substantial.<br />

5.2 CASE STUDY OF A GEOMORPHIC<br />

CONCEPTUAL MODEL<br />

5.2.1 Introduction<br />

The follow<strong>in</strong>g section describes <strong>the</strong> chronological development<br />

of <strong>the</strong> geomorphic conceptual model for <strong>the</strong><br />

restoration of <strong>the</strong> lower 32 km of <strong>the</strong> Snowy <strong>River</strong> <strong>in</strong><br />

south-eastern Australia. The Snowy <strong>River</strong> dra<strong>in</strong>s a catchment<br />

of 15 800 km2 situated <strong>in</strong> New South Wales and Victoria<br />

<strong>in</strong> south-east Australia, before discharg<strong>in</strong>g <strong>in</strong>to Bass<br />

Strait (Brizga and F<strong>in</strong>layson, 1994). For most of its course,<br />

<strong>the</strong> river fl ows <strong>in</strong> a narrow valley, that only widens around<br />

Bete Belong (Figure 5.1). The river below Bete Belong is<br />

a perched, sand-bed, channel between 70 m and 170 m<br />

wide. The downstream end of <strong>the</strong> Snowy <strong>River</strong> is estuar<strong>in</strong>e,<br />

with tidal <strong>in</strong>fl uences persist<strong>in</strong>g upstream to Orbost.<br />

Figure 5.1 The lower Snowy <strong>River</strong> <strong>in</strong> Victoria (from F<strong>in</strong>layson and Bird, 1989)<br />

When <strong>the</strong> fi rst European settlers arrived <strong>in</strong> <strong>the</strong> late<br />

1840s, <strong>the</strong> Snowy <strong>River</strong> fl oodpla<strong>in</strong> was swampy wetlands<br />

away from <strong>the</strong> channel, with <strong>the</strong> higher levees along <strong>the</strong><br />

channel covered <strong>in</strong> warm temperate ra<strong>in</strong>forest (described<br />

as ‘jungle’) (Owen, 1997). By <strong>the</strong> 1880s, <strong>the</strong> stream banks<br />

had been cleared and <strong>the</strong> clear<strong>in</strong>g and dra<strong>in</strong><strong>in</strong>g of <strong>the</strong><br />

wetlands had begun. By <strong>the</strong> 1930s, artifi cial fl ood levees<br />

had been built along much of <strong>the</strong> river and much of <strong>the</strong><br />

swampy wetlands had been dra<strong>in</strong>ed for graz<strong>in</strong>g. Over<br />

<strong>the</strong> same period, large woody debris was removed from<br />

<strong>the</strong> stream. This began <strong>in</strong> <strong>the</strong> 1880s to aid navigation<br />

(Seddon, 1994) and reached a peak <strong>in</strong> <strong>the</strong> 1950s to combat<br />

perceived aggradation of <strong>the</strong> bed and to reduce fl ood peaks<br />

(F<strong>in</strong>layson and Bird, 1989).<br />

A second phase of disturbance was <strong>in</strong>itiated by construction<br />

<strong>in</strong> <strong>the</strong> river’s headwaters under <strong>the</strong> Snowy Mounta<strong>in</strong><br />

Scheme. Completed <strong>in</strong> 1967, <strong>the</strong> Snowy Mounta<strong>in</strong> Scheme<br />

diverts water out of <strong>the</strong> Snowy <strong>River</strong> catchment for hydroelectric<br />

power generation and for <strong>the</strong> supply of water for<br />

irrigation <strong>in</strong> <strong>the</strong> neighbor<strong>in</strong>g Murray and Murrumbidgee


Conceptual and Ma<strong>the</strong>matical Modell<strong>in</strong>g <strong>in</strong> <strong>River</strong> <strong>Restoration</strong>: Do We Have Unreasonable Confi dence? 65<br />

catchments. Ersk<strong>in</strong>e et al. (1999) concluded that <strong>the</strong> Snowy<br />

Mounta<strong>in</strong> Scheme had reduced median and low fl ows by<br />

60–70%, as far downstream as Jarrahmond.<br />

5.2.2 The Geomorphic Model<br />

The follow<strong>in</strong>g is a chronological summary of <strong>the</strong> history of<br />

stream channel change recorded by various workers on <strong>the</strong><br />

lower Snowy <strong>River</strong> and explanations for <strong>the</strong>se changes.<br />

1. There are anecdotal suggestions that <strong>the</strong> lower Snowy<br />

<strong>River</strong> <strong>in</strong>creased its width at <strong>the</strong> end of <strong>the</strong> n<strong>in</strong>eteenth<br />

century. For example, <strong>the</strong> Snowy <strong>River</strong> Mail (July 1,<br />

1884) wrote:<br />

‘The river which, when settlement fi rst appeared<br />

here, was not half its present width, is <strong>in</strong>creas<strong>in</strong>g at<br />

every fl ood, widen<strong>in</strong>g and shallow<strong>in</strong>g <strong>the</strong> channel<br />

and precipitat<strong>in</strong>g huge trees <strong>in</strong>to its bed . . .’ (cited <strong>in</strong><br />

Seddon, 1994).<br />

In review<strong>in</strong>g <strong>the</strong>se claims, F<strong>in</strong>layson and Bird (1989)<br />

concluded that <strong>the</strong>re was not suffi cient evidence to substantiate<br />

this catastrophic <strong>in</strong>crease <strong>in</strong> width.<br />

2. S<strong>in</strong>ce <strong>the</strong> 1930s, people liv<strong>in</strong>g along <strong>the</strong> Snowy <strong>River</strong><br />

have been conv<strong>in</strong>ced that <strong>the</strong> river was becom<strong>in</strong>g<br />

shallower and fi ll<strong>in</strong>g with sand (Strom, 1936). Flush<strong>in</strong>g<br />

out this sand was probably <strong>the</strong> earliest ‘restoration’<br />

target for <strong>the</strong> river, lead<strong>in</strong>g to desnagg<strong>in</strong>g and o<strong>the</strong>r<br />

works <strong>in</strong> <strong>the</strong> stream <strong>in</strong> <strong>the</strong> 1950s. Despite <strong>the</strong> local<br />

conviction that <strong>the</strong> bed has aggraded, comparisons of<br />

16 repeat cross-sections, dat<strong>in</strong>g from <strong>the</strong> 1920s, along<br />

<strong>the</strong> Jarrahmond reach, do not support this view (Gippel,<br />

2002). Instead <strong>the</strong> bed level fl uctuates over a range of<br />

±2 m. Brizga and F<strong>in</strong>layson (1994) suggested that <strong>the</strong><br />

reduced fl ows of <strong>the</strong> river s<strong>in</strong>ce regulation mean that<br />

more of <strong>the</strong> bed is now visible, lead<strong>in</strong>g to <strong>the</strong> illusion<br />

of aggradation. This is a controversial suggestion<br />

amongst <strong>the</strong> local people, who still see ameliorat<strong>in</strong>g <strong>the</strong><br />

effects of sedimentation as <strong>the</strong> major restoration target<br />

for <strong>the</strong> river. This perception may come, <strong>in</strong> part, from<br />

<strong>the</strong> fact that some deep pools <strong>in</strong> <strong>the</strong> river have certa<strong>in</strong>ly<br />

fi lled-<strong>in</strong> s<strong>in</strong>ce <strong>the</strong> 1970s, particularly around Bete<br />

Bolong (Gippel, 2002).<br />

3. In <strong>the</strong> 1990s, attention turned to <strong>the</strong> effect of <strong>the</strong> Snowy<br />

Mounta<strong>in</strong> Scheme on <strong>the</strong> geomorphology of <strong>the</strong> river.<br />

Brizga and F<strong>in</strong>layson (1992) mentioned <strong>the</strong> loss of<br />

lateral bars, and <strong>the</strong>ir associated pools <strong>in</strong> <strong>the</strong> lower<br />

Snowy <strong>River</strong>, and speculated that <strong>the</strong> cause could be<br />

fl ow regulation (Figure 5.2). Ersk<strong>in</strong>e and Tilleard<br />

(1997) described <strong>the</strong> loss of <strong>the</strong> bars and related it more<br />

Figure 5.2 1940 aerial photograph of <strong>the</strong> Lower Snowy <strong>River</strong><br />

upstream of Lynn’s Gulch, show<strong>in</strong>g well developed lateral bars<br />

and associated pools<br />

strongly to <strong>the</strong> loss of some formative discharges after<br />

regulation <strong>in</strong> 1967. ‘. . . rhythmically spaced, bankattached,<br />

alternate side bars with well defi ned poolriffl<br />

e sequence were present above <strong>the</strong> estuary at Bete<br />

Bolong before 1967 and <strong>the</strong>y have never reformed<br />

s<strong>in</strong>ce <strong>the</strong>n . . .’ (Ersk<strong>in</strong>e et al., 1999). Stewardson<br />

(1998) concluded that, not only has <strong>the</strong> frequency of<br />

fl ows that are thought to form bars and pools fallen<br />

dramatically, but <strong>the</strong> <strong>in</strong>cidence of low fl ows that can<br />

potentially <strong>in</strong>-fi ll <strong>the</strong> pools with sediment has <strong>in</strong>creased.<br />

Replac<strong>in</strong>g deep pools with a ‘plane bed’ of sand is<br />

thought to provide poor habitat, particularly for migrat<strong>in</strong>g<br />

fi sh (Raadick and O’Connor, 1997). Twelve of <strong>the</strong><br />

seventeen fi sh species found <strong>in</strong> <strong>the</strong> river are migratory.<br />

Return<strong>in</strong>g <strong>the</strong> pools to <strong>the</strong> lower Snowy <strong>River</strong> presented<br />

an elegant and achievable goal for stream restoration<br />

and was <strong>the</strong> ma<strong>in</strong> recommendation of <strong>the</strong> fi rst<br />

restoration plan (ID&A, 1998). The recommendation


66 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

was also supported by <strong>the</strong> Snowy Water Inquiry<br />

(1998a,b). Follow<strong>in</strong>g <strong>the</strong>se reports, <strong>the</strong> effort <strong>in</strong> <strong>the</strong><br />

project swung to design<strong>in</strong>g timber pile fi elds (retards)<br />

at <strong>the</strong> historical locations of <strong>the</strong> bank attached side<br />

bars. It was hoped that <strong>the</strong>se retards would encourage<br />

sand deposition and eventually scour pools (ID&A,<br />

1998; Gippel et al., 2002).<br />

4. Gippel (2002) has comprehensively reviewed <strong>the</strong> evidence<br />

for channel change <strong>in</strong> <strong>the</strong> lower Snowy <strong>River</strong>,<br />

add<strong>in</strong>g <strong>the</strong> evidence from two student <strong>the</strong>ses. Gippel<br />

concludes that it is likely that <strong>the</strong> Snowy <strong>River</strong> did<br />

dramatically <strong>in</strong>crease <strong>in</strong> width after <strong>the</strong> 1870 fl ood, <strong>in</strong><br />

common with o<strong>the</strong>r rivers <strong>in</strong> Gippsland (Brooks and<br />

Brierley, 1997; Brooks et al., 2003). This suggests that<br />

<strong>the</strong> pre-1967 un-regulated river (that has up till now<br />

formed <strong>the</strong> ‘reference’ for <strong>the</strong> restoration strategy) was<br />

<strong>in</strong> far from ‘natural condition’. Gippel concludes that<br />

<strong>the</strong> evidence that regulation removed <strong>the</strong> naturally<br />

occurr<strong>in</strong>g lateral bars, is weak. An alternative possibility<br />

is that bars did not form until <strong>the</strong> river widened after<br />

disturbance, dur<strong>in</strong>g <strong>the</strong> early period of European settlement<br />

and, even <strong>the</strong>n, <strong>the</strong> bars occurred only when<br />

hydrological conditions were ideal. Regulation caused<br />

<strong>the</strong>se ideal hydrological conditions to be less likely. In<br />

addition, Gippel concluded that:<br />

After 1940, <strong>the</strong> alternate bar and pool morphology<br />

only ever existed <strong>in</strong> <strong>the</strong> straight Jarrahmond/Bete<br />

Bolong reach (10.2 km of <strong>the</strong> 32 km lowland section<br />

of <strong>the</strong> river).<br />

It has been <strong>in</strong>correctly assumed that <strong>the</strong> alternate bars<br />

always occurred <strong>in</strong> <strong>the</strong> same location, with <strong>the</strong> same<br />

wavelength, when <strong>in</strong> fact <strong>the</strong>y moved and changed<br />

form.<br />

Even more fundamentally, <strong>the</strong> l<strong>in</strong>k between pools<br />

and fi sh diversity has never been well established.<br />

Even <strong>in</strong> <strong>the</strong> orig<strong>in</strong>al report by Raadik and O’Connor<br />

(1997), <strong>the</strong> abundance and diversity of fi sh <strong>in</strong> <strong>the</strong><br />

reaches without pools were found to be higher than<br />

<strong>in</strong> <strong>the</strong> reference reaches with pools.<br />

There is some question about whe<strong>the</strong>r <strong>the</strong> pools can<br />

be susta<strong>in</strong>ed. The persistent low fl ows that characterise<br />

<strong>the</strong> regulated regime could quickly <strong>in</strong>-fi ll any<br />

pools that are scoured by favourable fl ow events.<br />

At present, lateral bars and pools rema<strong>in</strong> <strong>the</strong> major<br />

focus of <strong>the</strong> restoration plan on <strong>the</strong> Snowy <strong>River</strong>. The<br />

Victorian Government is develop<strong>in</strong>g a major trial (US$ 1.5<br />

million) of retard structures (<strong>in</strong> <strong>the</strong> fi eld and <strong>in</strong> fl umes), as<br />

a pilot project before build<strong>in</strong>g <strong>the</strong> major pile-fi elds. The<br />

trial has been held up because <strong>the</strong> local community would<br />

not accept <strong>the</strong> assurances of <strong>the</strong> eng<strong>in</strong>eers and geomor-<br />

phologists that <strong>the</strong> proposed log structures would not<br />

cause any change <strong>in</strong> fl ood stage or duration. Ironically, <strong>the</strong><br />

project has been stopped, not because of scientifi c uncerta<strong>in</strong>ty<br />

about <strong>the</strong> goals of <strong>the</strong> project, but because <strong>the</strong><br />

community rema<strong>in</strong>s uncerta<strong>in</strong> about someth<strong>in</strong>g that <strong>the</strong><br />

eng<strong>in</strong>eers and scientists are certa<strong>in</strong> of: <strong>the</strong> m<strong>in</strong>or hydraulic<br />

effect of <strong>the</strong> works on fl oods.<br />

5.2.3 Analysis of Uncerta<strong>in</strong>ties<br />

The foundation for stream restoration projects are conceptual<br />

models of physical and biological change. This case<br />

study demonstrates that <strong>the</strong> problem with much of <strong>the</strong><br />

proposed restoration comes from a false sense of certa<strong>in</strong>ty<br />

<strong>in</strong> <strong>the</strong>se models. Several geomorphologists have contributed<br />

to a conceptual model of channel change. The restoration<br />

plan has proceeded on <strong>the</strong> basis of a conceptual<br />

model that dismissed aggradation by sand as a geomorphic<br />

change, but did identify a clear co<strong>in</strong>cidence between<br />

fl ow regulation and <strong>the</strong> loss of lateral bars. This neat conceptual<br />

model <strong>the</strong>n formed <strong>the</strong> basis for a major restoration<br />

project <strong>in</strong>volv<strong>in</strong>g eng<strong>in</strong>eer<strong>in</strong>g works to artifi cially<br />

recreate lateral bars and pools. Over <strong>the</strong> last fi ve years, <strong>the</strong><br />

eng<strong>in</strong>eer<strong>in</strong>g aspects of <strong>the</strong> project have taken hold. The<br />

key question for managers now is not whe<strong>the</strong>r pools are<br />

an appropriate goal but which eng<strong>in</strong>eer<strong>in</strong>g design will<br />

develop pools most effi ciently.<br />

It is reasonable to imag<strong>in</strong>e that managers beg<strong>in</strong> by be<strong>in</strong>g<br />

uncerta<strong>in</strong> about how a geomorphic system functions. They<br />

<strong>the</strong>n commission <strong>in</strong>vestigations that lead to progressively<br />

greater certa<strong>in</strong>ty, until restoration decisions can be made<br />

with confi dence. In <strong>the</strong> turbulent boundary between<br />

science and management (Cullen, 1989), however, certa<strong>in</strong>ty<br />

is a fi ckle commodity. Consider <strong>the</strong> chronology of<br />

certa<strong>in</strong>ty <strong>in</strong> <strong>the</strong> Snowy <strong>River</strong> project:<br />

1. The local community and <strong>the</strong> river managers were<br />

completely certa<strong>in</strong> that <strong>the</strong> river was aggrad<strong>in</strong>g and that<br />

<strong>the</strong> appropriate restoration strategy was to somehow<br />

remove <strong>the</strong> sand. Theories of aggradation were subsequently<br />

dismissed follow<strong>in</strong>g <strong>the</strong> geomorphic <strong>in</strong>vestigation<br />

of Brizga and F<strong>in</strong>layson (1994).<br />

2. Early geomorphic assessments concluded that regulation<br />

had led to <strong>the</strong> loss of alternate bars and restoration<br />

plans were developed to restore <strong>the</strong>m. Gippel’s (2002)<br />

subsequent review revealed strong evidence of channel<br />

widen<strong>in</strong>g <strong>in</strong> <strong>the</strong> late 1800s. S<strong>in</strong>ce alternate bars have<br />

only been found <strong>in</strong> wider reaches of <strong>the</strong> lower Snowy<br />

<strong>River</strong>, and only dur<strong>in</strong>g ideal hydrological periods,<br />

alternate bars were probably not a natural feature of <strong>the</strong><br />

channel prior to European settlement.


Conceptual and Ma<strong>the</strong>matical Modell<strong>in</strong>g <strong>in</strong> <strong>River</strong> <strong>Restoration</strong>: Do We Have Unreasonable Confi dence? 67<br />

3. F<strong>in</strong>ally, <strong>the</strong> restoration trial is be<strong>in</strong>g delayed because<br />

scientists are not able to conv<strong>in</strong>ce local stakeholders<br />

about <strong>the</strong> m<strong>in</strong>imal hydraulic effect of <strong>the</strong> proposed<br />

works.<br />

There are many examples where geomorphic conceptual<br />

models have been found want<strong>in</strong>g when used as a basis<br />

for management. Much of <strong>the</strong> criticism <strong>in</strong> <strong>the</strong> literature<br />

<strong>in</strong>volves situations where ‘real geomorphology’ has been<br />

replaced by ‘cookbook’ approaches (Sear, 1994). Much of<br />

this criticism has been directed at restoration projects<br />

based on classifi cation systems, such as that of Rosgen<br />

(1996). In one example, Kondolf et al. (2001) described<br />

a meander<strong>in</strong>g channel <strong>in</strong>appropriately constructed <strong>in</strong> a<br />

naturally braided stream. Miller and Ritter (1996) provide<br />

a more general review of <strong>the</strong> Rosgen method. The Snowy<br />

<strong>River</strong> Project, by contrast, is not superfi cial geomorphology,<br />

but uses <strong>the</strong> earnest pr<strong>in</strong>ciples recommended by<br />

geomorphologists (Kondolf, 2000). Even ‘real geomorphology’<br />

can be uncerta<strong>in</strong>. Kondolf and Micheli (1995)<br />

were correct when <strong>the</strong>y argued that every stream restoration<br />

project must be considered an experiment. This sits<br />

well with <strong>the</strong> sometimes stark uncerta<strong>in</strong>ties that surround<br />

diagnosis of geomorphic ‘problems’. However, it does not<br />

sit well with multi-million dollar <strong>in</strong>tervention projects.<br />

The lower Snowy <strong>River</strong> restoration project, for example,<br />

is planned to cost US$20 million, and this is before <strong>the</strong><br />

environmental fl ow component is <strong>in</strong>cluded.<br />

5.2.4 Discussion<br />

Process based conceptual models of stream channel<br />

change (with or without biological conceptual models) are<br />

one of <strong>the</strong> key contributions of geomorphology to stream<br />

restoration. However, <strong>the</strong>y are also a major source of<br />

uncerta<strong>in</strong>ty. A restoration project will proceed on <strong>the</strong> basis<br />

of a geomorphic conceptual model about which everybody<br />

is <strong>in</strong>itially confi dent. However, <strong>the</strong> Snowy <strong>River</strong> case<br />

study shows that this confi dence can be mislead<strong>in</strong>g.<br />

Fur<strong>the</strong>r <strong>in</strong>vestigation can reduce confi dence <strong>in</strong> a model,<br />

but by <strong>the</strong> time geomorphologists have settled on a conceptual<br />

model that deserves confi dence, <strong>the</strong> restoration<br />

process has moved on. Chang<strong>in</strong>g <strong>the</strong> model becomes<br />

<strong>in</strong>creas<strong>in</strong>gly diffi cult <strong>in</strong> <strong>the</strong> political and management<br />

process. Multi-million dollar restoration projects can be<br />

launched on <strong>the</strong> basis of uncerta<strong>in</strong> conceptual models. Not<br />

surpris<strong>in</strong>gly, managers and eng<strong>in</strong>eers are impatient to<br />

proceed to <strong>the</strong> ‘real’ bus<strong>in</strong>ess of restoration, which is<br />

build<strong>in</strong>g th<strong>in</strong>gs and chang<strong>in</strong>g th<strong>in</strong>gs. At least here <strong>the</strong><br />

uncerta<strong>in</strong>ty can be quantifi ed.<br />

Geomorphologists have argued strongly <strong>the</strong> central<br />

importance of understand<strong>in</strong>g <strong>the</strong> geomorphic context of<br />

restoration projects to ensure <strong>the</strong>ir susta<strong>in</strong>ability. If this<br />

is <strong>the</strong> case, a restoration plan will be underm<strong>in</strong>ed if <strong>the</strong><br />

conceptual model from which it was developed is wrong.<br />

In <strong>the</strong> case of <strong>the</strong> Snowy <strong>River</strong>, <strong>the</strong> <strong>in</strong>itial model of aggradation<br />

led to calls for artifi cial sand extraction to restore<br />

<strong>the</strong> river. However, such efforts would have been unsusta<strong>in</strong>able<br />

s<strong>in</strong>ce <strong>the</strong> sediment <strong>in</strong> <strong>the</strong> Snowy <strong>River</strong> bed was<br />

not a discrete slug of sand. Any pool formed by sand<br />

extraction would have been <strong>in</strong>fi lled dur<strong>in</strong>g subsequent<br />

storms.<br />

The accuracy of a conceptual model can be critical to<br />

<strong>the</strong> success of a project. However, <strong>the</strong>re is often only a<br />

poor basis for judg<strong>in</strong>g <strong>the</strong> levels of uncerta<strong>in</strong>ty <strong>in</strong> those<br />

models. Some conceptual models are reasonably simple<br />

and have been tested <strong>in</strong> numerous situations. The channelised<br />

stream and sand-slug models are examples of welltested<br />

models <strong>in</strong> general terms, although apply<strong>in</strong>g <strong>the</strong>m <strong>in</strong><br />

specifi c cases is challeng<strong>in</strong>g because <strong>the</strong>y provide a direction<br />

of change ra<strong>the</strong>r than ei<strong>the</strong>r rates or magnitudes.<br />

So, given that each stream requires a variant of a conceptual<br />

model, how is <strong>the</strong> certa<strong>in</strong>ty of that model judged?<br />

At present <strong>the</strong> model will be presented, with more or less<br />

confi dence, after a geomorphic study. That confi dence is<br />

based on ei<strong>the</strong>r <strong>the</strong> weight of reconstructed evidence (‘all<br />

12 of <strong>the</strong> cross-sections changed <strong>in</strong> <strong>the</strong> same way’), on<br />

precedents provided by analogous cases (‘a very similar<br />

model has been described on three nearby rivers’), or on<br />

<strong>the</strong> reputation and forcefulness of <strong>the</strong> <strong>in</strong>vestigator. Given<br />

that most of <strong>the</strong>se restoration projects cannot wait for<br />

models to be published <strong>in</strong> <strong>the</strong> peer-reviewed literature,<br />

how can managers judge <strong>the</strong> uncerta<strong>in</strong>ty of <strong>the</strong>se conceptual<br />

models? Here are fi ve proposals:<br />

1. The most obvious way to improve confi dence is to<br />

subject a conceptual model to anonymous, external<br />

review, by dis<strong>in</strong>terested ‘experts’. Whilst this sounds<br />

easy, <strong>the</strong> pool of appropriate reviewers can be small<br />

and, from a manager’s perspective, plans can get bogged<br />

down <strong>in</strong> what appear to be petty, academic debates.<br />

2. Simple guidel<strong>in</strong>es could be established for evaluat<strong>in</strong>g<br />

<strong>the</strong> uncerta<strong>in</strong>ty of a conceptual model based on <strong>the</strong> type<br />

and strength of evidence provided. Evidence <strong>in</strong> <strong>the</strong><br />

form of a ma<strong>the</strong>matical expression of <strong>the</strong> conceptual<br />

model tested on long term geomorphic data should<br />

provide more confi dence <strong>in</strong> <strong>the</strong> model than a model<br />

based on anecdotal accounts. The guidel<strong>in</strong>es could be<br />

applied by <strong>the</strong> geomorphologists develop<strong>in</strong>g <strong>the</strong> model<br />

to advise managers on model uncerta<strong>in</strong>ty. It would<br />

also be possible to use <strong>the</strong> guidel<strong>in</strong>es to recommend<br />

additional <strong>in</strong>vestigations that could reduce model<br />

uncerta<strong>in</strong>ty. Alternate conceptual models could also be<br />

compared us<strong>in</strong>g <strong>the</strong>se guidel<strong>in</strong>es.


68 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

3. A rule-of-thumb <strong>in</strong> eng<strong>in</strong>eer<strong>in</strong>g projects is that about<br />

10% of <strong>the</strong> total cost of a project should be spent on<br />

design. Given that stream restoration projects tend to<br />

cost AUS$10 5 –$10 6 , perhaps a good guide is to argue<br />

that all of <strong>the</strong> design would be cost<strong>in</strong>g around $US<br />

10 000 to $US 100 000, with <strong>the</strong> conceptual model<br />

absorb<strong>in</strong>g from 20–40% of that amount. This might<br />

avoid <strong>the</strong> situation described <strong>in</strong> <strong>the</strong> Snowy <strong>River</strong><br />

Project, where a simple/elegant conceptual model<br />

based on a modest study sets <strong>in</strong> tra<strong>in</strong> a management<br />

response that it is hard to modify. A problem with this<br />

approach is that <strong>the</strong> geomorphic <strong>in</strong>vestigation that<br />

establishes <strong>the</strong> conceptual model may be commissioned<br />

before <strong>the</strong> full costs of <strong>the</strong> restoration project<br />

are known. In this case, additional geomorphic <strong>in</strong>vestigations<br />

may be commissioned later <strong>in</strong> <strong>the</strong> plann<strong>in</strong>g of<br />

<strong>the</strong> project after prelim<strong>in</strong>ary estimates of project costs.<br />

The challenge is to avoid advanc<strong>in</strong>g too far with <strong>the</strong><br />

design before <strong>the</strong> conceptual model is fi nalised.<br />

4. Modellers often develop <strong>the</strong> model with one set of data<br />

and test it on ano<strong>the</strong>r set. Thus, some data are held back<br />

for verifi cation. A similar approach could be used with<br />

conceptual models. For example, <strong>in</strong> <strong>the</strong> Snowy <strong>River</strong><br />

project, a proportion of aerial photographs could have<br />

been held back and used to verify <strong>the</strong> model later<br />

(given that <strong>the</strong>re are eight sets of photos, this may be<br />

possible).<br />

5. All of <strong>the</strong> data and <strong>in</strong>formation could be collected and<br />

collated and <strong>the</strong>n passed onto an appropriate third party<br />

(or more than one), without <strong>in</strong>terpretation, so that an<br />

<strong>in</strong>dependent <strong>in</strong>terpretation of <strong>the</strong> data could be developed.<br />

The diffi culty with this is that <strong>the</strong> conceptual<br />

model developed by <strong>the</strong> third party may require a different<br />

type of data for test<strong>in</strong>g, which would require<br />

fur<strong>the</strong>r data collation, add<strong>in</strong>g substantially to project<br />

costs.<br />

5.3 CASE STUDY OF A GEOMORPHIC<br />

DESIGN MODEL<br />

5.3.1 Introduction<br />

So far, <strong>the</strong> role of geomorphologists <strong>in</strong> us<strong>in</strong>g conceptual<br />

models to identify ‘reference’ states and actions that can<br />

move a stream toward <strong>the</strong> ‘reference’ state have been discussed.<br />

The conceptual model will help to identify a restoration<br />

action, but a ma<strong>the</strong>matical model can be required<br />

to design specifi c aspects of <strong>the</strong> <strong>in</strong>tervention. The most<br />

common restoration <strong>in</strong>terventions recommended <strong>in</strong> <strong>the</strong><br />

nor<strong>the</strong>rn hemisphere relate to structural and fl ow changes<br />

that will improve <strong>the</strong> success of fi sh populations. Hav<strong>in</strong>g<br />

been <strong>in</strong>volved <strong>in</strong> many such projects <strong>in</strong> Australia, we have<br />

noted <strong>the</strong> tendency for managers (and ecologists) to<br />

assume that <strong>the</strong> ‘physical stuff’ – <strong>the</strong> hydrology, hydraulics<br />

and geomorphology – is well understood, with low errors<br />

and low uncerta<strong>in</strong>ty. They assume that most uncerta<strong>in</strong>ty<br />

lies <strong>in</strong> <strong>the</strong> biological responses to <strong>in</strong>tervention. The follow<strong>in</strong>g<br />

section suggests that <strong>the</strong>y may be giv<strong>in</strong>g <strong>the</strong> geomorphological<br />

studies we exam<strong>in</strong>e too much credit!<br />

In this section we exam<strong>in</strong>e <strong>the</strong> uncerta<strong>in</strong>ties <strong>in</strong> a ma<strong>the</strong>matical<br />

model used to design a restoration action. The<br />

case study concerns <strong>the</strong> design of fl ush<strong>in</strong>g-fl ows for a<br />

gravel-bed stream. Numerous authors have identifi ed <strong>the</strong><br />

problem of fi ne sediment <strong>in</strong>fi ltration <strong>in</strong>to a stream bed<br />

known as colmation (Sear, 1993; O’Neill and Kuhns,<br />

1994; Milhous, 1995; Kondolf and Wilcock, 1996). The<br />

loss of competent fl ows below dams means that coarse<br />

beds are <strong>in</strong>fi ltrated by fi ne sediment, damag<strong>in</strong>g habitat for<br />

macro<strong>in</strong>vertebrates and fi sh. The only way to clean out <strong>the</strong><br />

bed is to <strong>in</strong>itiate movement of <strong>the</strong> coarse fraction. A<br />

common geomorphic goal of fl ow management is to ‘turnover<br />

<strong>the</strong> bed’. The uncerta<strong>in</strong>ty <strong>in</strong> this question is whe<strong>the</strong>r<br />

<strong>the</strong> predicted fl ow will turn-over <strong>the</strong> bed. The cost of <strong>the</strong><br />

uncerta<strong>in</strong>ty is <strong>the</strong> chance that <strong>the</strong> bed does not move when<br />

<strong>the</strong> target amount of water is released. If <strong>the</strong> bed is not<br />

turned over, <strong>the</strong>n that volume of water has been wasted.<br />

Similarly, if <strong>the</strong> bed moves at a discharge below <strong>the</strong> target<br />

discharge, <strong>the</strong>n <strong>the</strong> extra volume of water released has<br />

been wasted. Thus, <strong>in</strong> this example, <strong>the</strong> uncerta<strong>in</strong>ty can<br />

be expressed <strong>in</strong> terms of <strong>the</strong> extra water that has to be<br />

released to be confi dent that <strong>the</strong> bed will fl ush. This also<br />

allows <strong>the</strong> relative sav<strong>in</strong>g <strong>in</strong> water to be estimated if managers<br />

do various th<strong>in</strong>gs to reduce <strong>the</strong> uncerta<strong>in</strong>ty.<br />

The case study considered is that of fl ush<strong>in</strong>g fl ows for<br />

<strong>the</strong> mid-Goulburn <strong>River</strong>, downstream of <strong>the</strong> Eildon Dam,<br />

<strong>in</strong> nor<strong>the</strong>rn Victoria, Australia. The Goulburn <strong>River</strong> is <strong>the</strong><br />

largest stream <strong>in</strong> Victoria (catchment of 20 000 km 2 ) and<br />

<strong>the</strong> largest Victorian tributary to <strong>the</strong> Murray <strong>River</strong>. It is a<br />

meander<strong>in</strong>g, anabranch<strong>in</strong>g river, that is regulated for irrigation<br />

by <strong>the</strong> Eildon Dam. The bankfull discharge is not<br />

clearly defi ned but fl ow along anabranch channels beg<strong>in</strong>s<br />

at about 120 m 3 /s <strong>in</strong> <strong>the</strong> reach below <strong>the</strong> Eildon Dam.<br />

Abundant data is available for <strong>the</strong> stream, compiled for a<br />

recent environmental fl ow study (DNRE, 2002).<br />

5.3.2 The Geomorphic Model<br />

The volume of water required to deliver <strong>the</strong> fl ush<strong>in</strong>g fl ow<br />

is estimated <strong>in</strong> <strong>the</strong> follow<strong>in</strong>g three stages. It must be<br />

emphasised that this is an approach used for design<strong>in</strong>g a<br />

fl ush<strong>in</strong>g fl ow. This approach is not advocated for use <strong>in</strong><br />

o<strong>the</strong>r studies. Alternate methods for calculat<strong>in</strong>g fl ush<strong>in</strong>g<br />

fl ows might be considered, particulary given <strong>the</strong> magni-


Conceptual and Ma<strong>the</strong>matical Modell<strong>in</strong>g <strong>in</strong> <strong>River</strong> <strong>Restoration</strong>: Do We Have Unreasonable Confi dence? 69<br />

tude of uncerta<strong>in</strong>ties <strong>in</strong> this analysis described later. The<br />

three stages are:<br />

1. Estimat<strong>in</strong>g <strong>the</strong> critical bed shear stress for <strong>in</strong>cipient<br />

motion of <strong>the</strong> bed sediments us<strong>in</strong>g Shields’ entra<strong>in</strong>ment<br />

function; <strong>the</strong> critical shear stress is given by:<br />

τ crit = (γs − γw)d50τ* (5.1)<br />

where γ s and γw are <strong>the</strong> specifi c weights of <strong>the</strong> bed sediment<br />

and water respectively, d 50 is <strong>the</strong> median gra<strong>in</strong> size<br />

of <strong>the</strong> bed sediments and τ* is Shields’ dimensionless<br />

shear stress, often estimated as 0.045 for gravel bed<br />

rivers (Buffi ngton and Montgomery, 1997).<br />

2. Estimat<strong>in</strong>g <strong>the</strong> discharge at which <strong>the</strong> mean shear stress<br />

for <strong>the</strong> reach is equal to <strong>the</strong> critical shear stress for<br />

<strong>in</strong>cipient motion, us<strong>in</strong>g a one-dimensional hydraulic<br />

model (applied over a 2 km reach of <strong>the</strong> Goulburn<br />

<strong>River</strong>); and<br />

3. Estimat<strong>in</strong>g <strong>the</strong> volume of water required to mimic <strong>the</strong><br />

natural duration and frequency of fl ow spells dur<strong>in</strong>g<br />

which bed sediments are mobilised based on an analysis<br />

of a modelled natural fl ow series.<br />

Importantly, a one-dimensional hydraulic model is<br />

used ra<strong>the</strong>r than <strong>the</strong> more sophisticated two- or threedimensional<br />

models now available because it is <strong>the</strong><br />

approach widely used <strong>in</strong> practice for restoration projects.<br />

The uncerta<strong>in</strong>ties associated with a two-dimensional<br />

model will be different, but it is not possible to suggest<br />

whe<strong>the</strong>r <strong>the</strong>y would be smaller or larger than those us<strong>in</strong>g<br />

a one-dimensional analysis without a proper <strong>in</strong>vestigation.<br />

The modell<strong>in</strong>g procedure also <strong>in</strong>cludes <strong>the</strong> preparation of<br />

<strong>in</strong>put data which <strong>in</strong>cludes fi eld surveys and selection of<br />

an appropriate value for Shields’ dimensionless shear<br />

stress (Table 5.1). 0.045 is used as <strong>the</strong> value for Shields’<br />

dimensionless shear stress <strong>in</strong> <strong>the</strong> <strong>in</strong>itial model, but <strong>the</strong><br />

implications of uncerta<strong>in</strong>ty <strong>in</strong> this parameter are also considered.<br />

In this case study <strong>the</strong> concern is not so much with<br />

<strong>the</strong> correct value of this parameter as <strong>the</strong> effect of parameter<br />

uncerta<strong>in</strong>ty.<br />

Us<strong>in</strong>g a standard Wolman count (sample size = 100),<br />

<strong>the</strong> median bed material sediment size is estimated as<br />

30 mm. Us<strong>in</strong>g Shields’ entra<strong>in</strong>ment function, <strong>in</strong>cipient<br />

motion for <strong>the</strong> gra<strong>in</strong> size occurs at 22 N/m 2 . Seventeen<br />

cross-sections are surveyed along a 2 km reach of <strong>the</strong> mid-<br />

Goulburn river. A one-dimensional hydraulic model is<br />

calibrated for <strong>the</strong>se cross-sections us<strong>in</strong>g water levels surveyed<br />

at a fl ow close to <strong>the</strong> mean. Calibration was achieved<br />

by adjust<strong>in</strong>g <strong>the</strong> Mann<strong>in</strong>g n roughness parameter. Assum<strong>in</strong>g<br />

Mann<strong>in</strong>g n is <strong>in</strong>variant with discharge, <strong>the</strong> hydraulic<br />

model estimates that <strong>the</strong> mean shear stress for <strong>the</strong> reach<br />

is equal to 22 N/m 2 at a fl ow of 360 m 3 /s.<br />

To estimate <strong>the</strong> natural frequency of bed mobilisation<br />

events <strong>the</strong> natural fl ow regime is needed, but <strong>the</strong>re is no<br />

record of <strong>the</strong> natural fl ow regime <strong>in</strong> <strong>the</strong> mid-Goulburn<br />

<strong>River</strong> because it has been regulated s<strong>in</strong>ce fl ow gaug<strong>in</strong>g<br />

commenced. Instead a natural fl ow series is modelled<br />

us<strong>in</strong>g available streamfl ow data upstream of Eildon Dam.<br />

This requires estimation of fl ows from ungauged portions<br />

of <strong>the</strong> catchment by scal<strong>in</strong>g fl ows <strong>in</strong> <strong>the</strong> gauged catchments,<br />

comb<strong>in</strong><strong>in</strong>g fl ows from <strong>the</strong> various sub-catchments<br />

and rout<strong>in</strong>g fl ows to <strong>the</strong> study reach. Us<strong>in</strong>g this modelled<br />

natural fl ow series, it is estimated that 157 GL/year is<br />

required to mimic <strong>the</strong> natural frequency and duration of<br />

fl ows exceed<strong>in</strong>g 360 m 3 /s.<br />

5.3.3 Analysis of Uncerta<strong>in</strong>ties<br />

There are a number of sources of uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> estimated<br />

channel fl ush<strong>in</strong>g discharge and volume of water<br />

required to mimic <strong>the</strong> natural frequency and duration of<br />

channel fl ush<strong>in</strong>g events (Table 5.1). In this study, uncerta<strong>in</strong>ty<br />

<strong>in</strong> <strong>the</strong> estimated threshold discharge for sediment<br />

fl ush<strong>in</strong>g is assessed based on a consideration of:<br />

Table 5.1 Summary of errors and uncerta<strong>in</strong>ties considered <strong>in</strong> this analysis (<strong>the</strong> text <strong>in</strong> italics shows <strong>the</strong> values used <strong>in</strong> <strong>the</strong><br />

analysis)<br />

Source of Error Channel Hydraulics Critical Shear Flow Regime<br />

Sample uncerta<strong>in</strong>ty Cross-section sampl<strong>in</strong>g Number of particles <strong>in</strong> sample Number of years <strong>in</strong> <strong>the</strong> record<br />

(used n = 17)<br />

(used n = 100)<br />

(used n = 25)<br />

Measurement error (Neglected here) Measurement of gra<strong>in</strong><br />

diameter (± 5 mm)<br />

Error <strong>in</strong> rat<strong>in</strong>g curve (r2 = 0.95)<br />

Model error Mann<strong>in</strong>g n (Based on data Shields entra<strong>in</strong>ment function* Estimat<strong>in</strong>g fl ow from ungauged<br />

presented <strong>in</strong> Hicks and<br />

catchments (20% of catchment<br />

Mason, 1998)<br />

area) and fl ow rout<strong>in</strong>g*<br />

* see text for explanation.


70 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

• errors <strong>in</strong> measurement of gra<strong>in</strong> sizes;<br />

• sampl<strong>in</strong>g bed particles to estimate median gra<strong>in</strong> size;<br />

• estimat<strong>in</strong>g Shields’ dimensionless shear stress (τ*).<br />

• assum<strong>in</strong>g <strong>the</strong> channel roughness parameter (i.e. Mann<strong>in</strong>g<br />

n) is constant with discharge;<br />

• sampl<strong>in</strong>g of cross-sections from <strong>the</strong> 2 km reach.<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>the</strong> volume of water to mimic <strong>the</strong> duration<br />

and frequency of sediment fl ush<strong>in</strong>g fl ows is assessed based<br />

on a consideration of <strong>the</strong>se same factors <strong>in</strong> addition to:<br />

• errors <strong>in</strong> measur<strong>in</strong>g discharge us<strong>in</strong>g a rat<strong>in</strong>g curve at<br />

each gauge;<br />

• us<strong>in</strong>g a sample of <strong>the</strong> fl ow record to represent <strong>the</strong> long<br />

term fl ow regime;<br />

• errors <strong>in</strong> modell<strong>in</strong>g natural fl ows at <strong>the</strong> survey site.<br />

In this section <strong>the</strong> effect of <strong>the</strong>se uncerta<strong>in</strong>ties is quantifi<br />

ed us<strong>in</strong>g a Monte Carlo Analysis, which <strong>in</strong>volves runn<strong>in</strong>g<br />

<strong>the</strong> fl ush<strong>in</strong>g fl ow model many times (1000 <strong>in</strong> this case)<br />

with different, but equally plausible sets of <strong>in</strong>put parameters<br />

to generate a range of plausible model outputs (Manly,<br />

1997). For each of <strong>the</strong> 1000 replicates values were randomly<br />

chosen for each of <strong>the</strong> <strong>in</strong>put parameters from <strong>the</strong><br />

range of possible values. These were selected from a probability<br />

distribution centred on <strong>the</strong> best estimate used <strong>in</strong> <strong>the</strong><br />

orig<strong>in</strong>al analysis (described <strong>in</strong> <strong>the</strong> previous section). The<br />

comb<strong>in</strong>ed uncerta<strong>in</strong>ty is estimated by comb<strong>in</strong><strong>in</strong>g random<br />

selections of values for each <strong>in</strong>put variable.<br />

It is necessary to evaluate <strong>the</strong> uncerta<strong>in</strong>ty for each<br />

<strong>in</strong>put parameter so that it is known how it should be<br />

varied <strong>in</strong> <strong>the</strong> replicate model runs. A highly uncerta<strong>in</strong><br />

parameter should be varied over a bigger range than a more<br />

accurately known parameter. The methods used to evaluate<br />

parameter uncerta<strong>in</strong>ties are described below. In some<br />

cases, <strong>the</strong>se methods are obvious. For <strong>the</strong> case of <strong>the</strong><br />

median particle size, it is relatively straight forward to<br />

express <strong>the</strong> estimate of <strong>the</strong> median gra<strong>in</strong> size as a distribution<br />

of possible values, based on <strong>the</strong> size of <strong>the</strong> sample of<br />

bed gra<strong>in</strong>s. It is not so straightforward to evaluate uncerta<strong>in</strong>ties<br />

<strong>in</strong> recorded discharge series or calibrated values<br />

of <strong>the</strong> Mann<strong>in</strong>g channel roughness parameter (n). A best<br />

effort is made to quantify <strong>the</strong>se uncerta<strong>in</strong>ties, but <strong>the</strong>se are<br />

best described as models of uncerta<strong>in</strong>ty, albeit models<br />

which cannot be truly tested. The result is a stochastic<br />

model where some components are determ<strong>in</strong>istic and some<br />

are random.<br />

Median Particle Size<br />

The median gra<strong>in</strong> size is estimated from a sample of 100<br />

bed particles (a standard Wolman count). To estimate <strong>the</strong><br />

uncerta<strong>in</strong>ty associated with sampl<strong>in</strong>g gra<strong>in</strong> sizes, 1000<br />

replicate samples (each with 100 gra<strong>in</strong> sizes) are syn<strong>the</strong>tically<br />

generated by assum<strong>in</strong>g that <strong>the</strong> natural log of<br />

gra<strong>in</strong> sizes (<strong>in</strong> millimetres) are distributed normally with<br />

a mean of 3.4 and standard deviation of 0.51. This gives<br />

a true median particle size of 30 mm, although sample<br />

medians for each replicate will vary about this value.<br />

Errors <strong>in</strong> measurement of gra<strong>in</strong> size diameter are modelled<br />

as normally distributed with a mean of zero and standard<br />

deviation of 3 mm. This gives 90% confi dence <strong>in</strong>tervals<br />

on gra<strong>in</strong> size measurements of ±5 mm. To represent<br />

measurement errors, each gra<strong>in</strong> size <strong>in</strong> each of <strong>the</strong> 1000<br />

replicate samples is perturbed by add<strong>in</strong>g this random<br />

component.<br />

Entra<strong>in</strong>ment Threshold<br />

Buffi ngton and Montgomery (1997) collated values provided<br />

<strong>in</strong> <strong>the</strong> literature for <strong>the</strong> Shields’ dimensionless shear<br />

stress value for <strong>in</strong>cipient motion. A value of 0.045 is often<br />

used as <strong>the</strong> best estimate of τ* for gravel-bed rivers. The<br />

uncerta<strong>in</strong>ty <strong>in</strong> τ* is estimated from <strong>the</strong> range of values<br />

compiled by Buffi ngton and Montgomery that are likely<br />

to occur <strong>in</strong> gravel-bed streams with Reynolds roughness<br />

number <strong>in</strong> <strong>the</strong> range 25 to 1000 (Figure 5.3). The distribution<br />

of residuals for <strong>the</strong> regression <strong>in</strong> Figure 5.3 was replicated<br />

by assum<strong>in</strong>g that <strong>the</strong> distribution was log-normally<br />

distributed with a standard deviation <strong>in</strong> <strong>the</strong> log-error of<br />

0.03, and randomly generat<strong>in</strong>g 1000 replicate values us<strong>in</strong>g<br />

this error model.<br />

Dimensionless shear stress<br />

0.1<br />

y = 0.017x 0.145<br />

R 2 = 0.150<br />

n = 98<br />

0.01<br />

10 100 1000<br />

Reynolds roughness number<br />

Figure 5.3 Values of Reynolds roughness number (for <strong>the</strong> range<br />

25 to 1000) and dimensionless shear stress obta<strong>in</strong>ed from published<br />

<strong>in</strong>cipient motion studies and collated by Buffi ngton and<br />

Montgomery (1997)


Conceptual and Ma<strong>the</strong>matical Modell<strong>in</strong>g <strong>in</strong> <strong>River</strong> <strong>Restoration</strong>: Do We Have Unreasonable Confi dence? 71<br />

Survey Errors<br />

<strong>Uncerta<strong>in</strong>ty</strong> associated with cross-section survey errors<br />

were neglected <strong>in</strong> this study and likely to be small given<br />

<strong>the</strong> high accuracy of survey equipment relative to o<strong>the</strong>r<br />

sources of uncerta<strong>in</strong>ty. The survey of <strong>the</strong> Goulburn <strong>River</strong><br />

reach <strong>in</strong>cluded a sample of 17 evenly-spaced crosssections<br />

over a reach length of 2 km. There is uncerta<strong>in</strong>ty<br />

<strong>in</strong> <strong>the</strong> estimate of reach-averaged bed shear stress depend<strong>in</strong>g<br />

on how well <strong>the</strong> sampled cross-sections represent <strong>the</strong><br />

variability of <strong>the</strong> reach. To estimate this uncerta<strong>in</strong>ty, replicate<br />

samples of 17 cross-sections were generated us<strong>in</strong>g<br />

a bootstrap procedure (Manly, 1997) <strong>in</strong> which 17 crosssections<br />

were randomly selected from <strong>the</strong> 17 available,<br />

‘with replacement’. This bootstrap procedure was used to<br />

generate 1000 replicate samples.<br />

Errors <strong>in</strong> <strong>the</strong> Hydraulic Model<br />

The critical shear stress was converted <strong>in</strong>to a discharge<br />

us<strong>in</strong>g a one-dimensional hydraulic model. The major<br />

uncerta<strong>in</strong>ty <strong>in</strong> this method is <strong>the</strong> roughness coeffi cient. In<br />

<strong>the</strong> model, Mann<strong>in</strong>g n was calculated from a s<strong>in</strong>gle stage<br />

discharge measurement <strong>in</strong> <strong>the</strong> reach. The hydraulic model<strong>in</strong>g<br />

required <strong>the</strong> assumption that Mann<strong>in</strong>g n was <strong>the</strong>n<br />

constant over <strong>the</strong> range of discharges considered (up to<br />

bankfull). It is well known that Mann<strong>in</strong>g n varies with<br />

discharge, sometimes <strong>in</strong>creas<strong>in</strong>g and sometimes decreas<strong>in</strong>g<br />

with <strong>in</strong>creas<strong>in</strong>g fl ow. The error associated with this<br />

assumption was estimated us<strong>in</strong>g <strong>the</strong> large range of roughness<br />

estimates for gravel-bed New Zealand rivers (Hicks<br />

and Mason, 1998). These data reveal that <strong>the</strong> <strong>in</strong>verse of<br />

Mann<strong>in</strong>g roughness parameter ( 1 – n) is approximately proportional<br />

to <strong>the</strong> log of discharge <strong>in</strong> most cases (Stewardson<br />

and Anderson, 2002). A log regression was fi tted to data<br />

for each of 72 sites [provided by Hicks and Mason, 1998]<br />

to provide a constant (c) and coeffi cient (k) <strong>in</strong> this regression<br />

equation for each site:<br />

1<br />

c 1 k<br />

n<br />

Q ⎛ ⎞<br />

= ⎜ + ln − ⎟<br />

(5.2)<br />

⎝ Q ⎠<br />

where Q – is <strong>the</strong> mean daily fl ow at <strong>the</strong> site. About one<br />

quarter of <strong>the</strong> coeffi cients (k) were negative, <strong>in</strong>dicat<strong>in</strong>g<br />

<strong>in</strong>creas<strong>in</strong>g Mann<strong>in</strong>g n with discharge at <strong>the</strong>se sites. The<br />

parameter k defi nes variation <strong>in</strong> Mann<strong>in</strong>g n with discharge.<br />

For our site, c was determ<strong>in</strong>ed by calibration to <strong>the</strong><br />

observed water surface profi le (us<strong>in</strong>g k = 0). The change<br />

<strong>in</strong> shear stress with discharge was <strong>the</strong>n modelled us<strong>in</strong>g <strong>the</strong><br />

one-dimensional model, and run 1000 times, each time<br />

us<strong>in</strong>g a different value of k selected at random from <strong>the</strong><br />

72 values for <strong>the</strong> New Zealand rivers.<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>the</strong> Hydrology<br />

Flows at <strong>the</strong> Goulburn <strong>River</strong> site are regulated by operation<br />

of Lake Eildon, a large water supply reservoir,<br />

upstream of <strong>the</strong> site. Flow data from gaug<strong>in</strong>g stations<br />

located on n<strong>in</strong>e unregulated tributaries upstream of <strong>the</strong><br />

Goulburn <strong>River</strong> site were used to model natural daily fl ows<br />

at this site (i.e. fl ows that would occur <strong>in</strong> <strong>the</strong> absence of<br />

Lake Eildon, neglect<strong>in</strong>g rout<strong>in</strong>g effects). A 25 year series<br />

of daily fl ows was generated for <strong>the</strong> survey site from <strong>the</strong><br />

25 year fl ow records at n<strong>in</strong>e streamfl ow gauges. These<br />

25 year series are a sample of <strong>the</strong> long term fl ow regime.<br />

A bootstrap procedure was used to generate 1000 replicate<br />

25-year series by sampl<strong>in</strong>g random years from <strong>the</strong> 25 year<br />

sequences. Variability <strong>in</strong> <strong>the</strong> replicates represents uncerta<strong>in</strong>ty<br />

<strong>in</strong> <strong>the</strong> long term fl ow regime. This approach assumes<br />

fl ow <strong>in</strong>dependence between years.<br />

Flow modell<strong>in</strong>g for <strong>the</strong> study was based on records of<br />

discharge at n<strong>in</strong>e streamfl ow gauges on unregulated tributaries<br />

of <strong>the</strong> Goulburn <strong>River</strong>. Discharge is estimated at<br />

<strong>the</strong> gauges from rat<strong>in</strong>g curves. Rat<strong>in</strong>g curves are fi tted to<br />

periodic measurements of discharge and stage. There are<br />

errors <strong>in</strong> discharges provided by <strong>the</strong> rat<strong>in</strong>g curve as a<br />

consequence of errors <strong>in</strong> <strong>the</strong>se periodic measurements of<br />

stage and discharge. Rat<strong>in</strong>g curves are often fi tted by a<br />

log–log regression and an r2 of 0.95 is generally regarded<br />

as a good fi t (Clarke, 1999). Error <strong>in</strong> discharges derived<br />

from <strong>the</strong> rat<strong>in</strong>g curves are represented by a random component<br />

added to <strong>the</strong> log-discharge where <strong>the</strong> error is distributed<br />

normally with standard deviation of 0.05 times<br />

<strong>the</strong> mean of log-discharges recorded at <strong>the</strong> gauge. This can<br />

be represented by:<br />

Q′= e<br />

ln Q+ N(<br />

0005) ∑ ln Q<br />

n<br />

1<br />

, .<br />

(5.3)<br />

where Q is <strong>the</strong> recorded daily discharge, N(0,0.05) denotes<br />

a normally distributed random variable with mean of zero<br />

and standard deviation of 0.05, and n is <strong>the</strong> number of days<br />

of recorded fl ow. This error would result <strong>in</strong> an r 2 of 0.95<br />

for a rat<strong>in</strong>g curve fi tted by log–log regression.<br />

The 25 year natural daily fl ow series at <strong>the</strong> study site<br />

was modeled <strong>in</strong> two parts: (i) estimat<strong>in</strong>g fl ows from <strong>the</strong><br />

ungauged portion of <strong>the</strong> catchment by scal<strong>in</strong>g gauged<br />

fl ows <strong>in</strong> unregulated tributaries; and (ii) summ<strong>in</strong>g fl ows<br />

recorded at <strong>the</strong> streamfl ow gauges and estimated for <strong>the</strong><br />

ungauged portions of <strong>the</strong> catchment. Rout<strong>in</strong>g effects,<br />

which generally result <strong>in</strong> attenuation and delay of fl ood<br />

pulses fur<strong>the</strong>r downstream, were ignored because no data<br />

were available to calibrate a rout<strong>in</strong>g model. <strong>Uncerta<strong>in</strong>ty</strong><br />

<strong>in</strong> <strong>the</strong> scal<strong>in</strong>g parameter for <strong>the</strong> ungauged catchment<br />

was evaluated us<strong>in</strong>g a comparison of fl ows recorded at<br />

seven different gauges (see Appendix 5.1). There is some


72 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> modelled natural fl ows as a consequence<br />

of ignor<strong>in</strong>g rout<strong>in</strong>g effects. This uncerta<strong>in</strong>ty was represented<br />

us<strong>in</strong>g a rout<strong>in</strong>g model (described by Stewardson<br />

and Cott<strong>in</strong>gham, 2002), where <strong>the</strong> rout<strong>in</strong>g parameters<br />

were randomly perturbed accord<strong>in</strong>g to an assumed distribution<br />

of possible values (see Appendix 5.1).<br />

Results of Monte Carlo Analysis<br />

The critical shear stress for <strong>in</strong>cipient motion for this reach<br />

has wide confi dence limits vary<strong>in</strong>g from 18 N/m2 , up to<br />

26 N/m2 (Figure 5.4). A larger uncerta<strong>in</strong>ty applies to <strong>the</strong><br />

Shear stress (N/m 2 )<br />

Flow volume (GL/year)<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

800<br />

600<br />

400<br />

200<br />

reach shear stress<br />

critical shear stress<br />

0 200 400 600 800<br />

Flow (m 3 /s)<br />

(a)<br />

0<br />

0 200 400 600 800<br />

Flow threshold (m3 /s)<br />

(b)<br />

Figure 5.4 (a) reach-averaged shear stress estimated us<strong>in</strong>g a<br />

one-dimensional hydraulic model and <strong>the</strong> critical shear stress<br />

estimated us<strong>in</strong>g Shields, entra<strong>in</strong>ment function; and (b) <strong>the</strong> volume<br />

of artifi cial fl ow spells required to mimic <strong>the</strong> frequency and duration<br />

of natural fl ow spells for vary<strong>in</strong>g fl ow threshold. (Dashed<br />

l<strong>in</strong>es <strong>in</strong>dicate 90% confi dence <strong>in</strong>tervals based on consideration of<br />

all errors described <strong>in</strong> Table 5.1.)<br />

possible range of shear stresses that occur <strong>in</strong> <strong>the</strong> reach for<br />

a given discharge. The fl ush<strong>in</strong>g fl ow is estimated to be<br />

360 m 3 /s, but this has 90% confi dence limits of between<br />

250 m 3 /s and well over 500 m 3 /s.<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>the</strong> critical discharge for bed fl ush<strong>in</strong>g is<br />

reported us<strong>in</strong>g <strong>the</strong> magnitude of <strong>the</strong> 90% confi dence <strong>in</strong>terval<br />

(Table 5.2). Comb<strong>in</strong><strong>in</strong>g all sources of uncerta<strong>in</strong>ty, <strong>the</strong><br />

magnitude of this confi dence <strong>in</strong>terval is 560 m 3 /s. The<br />

major source of uncerta<strong>in</strong>ty <strong>in</strong> estimat<strong>in</strong>g <strong>the</strong> critical discharge<br />

comes from error <strong>in</strong> <strong>the</strong> estimate of Mann<strong>in</strong>g n <strong>in</strong><br />

<strong>the</strong> hydraulic model. This is because n is calibrated to a<br />

s<strong>in</strong>gle discharge (effectively <strong>the</strong> mean discharge of 60 m 3 /<br />

s) and becomes <strong>in</strong>creas<strong>in</strong>gly uncerta<strong>in</strong> as discharge<br />

<strong>in</strong>creases. This provides a massive 430 m 3 /s range of discharge<br />

(Table 5.2). By comparison, if <strong>the</strong> only source of<br />

uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> estimate was error <strong>in</strong> measur<strong>in</strong>g <strong>the</strong><br />

particle size, <strong>the</strong>n <strong>the</strong> range <strong>in</strong> <strong>the</strong> estimate would only be<br />

50 m 3 /s. The implication of Table 5.2 is that ever more<br />

detailed estimates of <strong>the</strong> bed-particle size distribution, or<br />

even <strong>in</strong>creas<strong>in</strong>gly sophisticated bed-load transport threshold<br />

models, will not remove <strong>the</strong> major uncerta<strong>in</strong>ty associated<br />

with <strong>the</strong> channel hydraulics.<br />

In a good environmental fl ow project, a manager would<br />

not simply specify a fl ow magnitude, <strong>the</strong> frequency and<br />

duration of that critical fl ow would also be specifi ed. A<br />

popular approach used to defi ne <strong>the</strong> frequency and duration<br />

is <strong>the</strong> ‘natural fl ow paradigm’ (Poff et al., 1997;<br />

Richter et al., 1997), which suggests that an environmental<br />

fl ow should mimic <strong>the</strong> natural regime. To estimate this <strong>the</strong><br />

average duration and frequency of critical transport events<br />

<strong>in</strong> <strong>the</strong> natural regime is calculated. The volume of water<br />

required to mimic <strong>the</strong>se events is <strong>the</strong> average duration<br />

multiplied by <strong>the</strong> frequency, multiplied by <strong>the</strong> threshold<br />

discharge. The curve <strong>in</strong> Figure 5.4(b) shows that, as <strong>the</strong><br />

fl ow threshold <strong>in</strong>creases, <strong>the</strong> total fl ow required decreases<br />

substantially. If, for example, <strong>the</strong> threshold fl ush<strong>in</strong>g fl ow<br />

Table 5.2 Size of 90% confi dence <strong>in</strong>terval (expressed <strong>in</strong> m 3 /s)<br />

for threshold discharge for <strong>in</strong>cipient motion for different<br />

sources of error (best estimate of threshold discharge is<br />

360 m 3 /s)<br />

Source of error Channel Hydraulics Critical Shear<br />

Measurement error Neglected 50<br />

Model error 430 180<br />

Sample error 200 120<br />

Comb<strong>in</strong>ed 540 220<br />

All sources comb<strong>in</strong>ed 560<br />

Note: The numbers are not additive. The comb<strong>in</strong>ed uncerta<strong>in</strong>ty comes<br />

from MonteCarlo simulations us<strong>in</strong>g <strong>the</strong> full distributions. This table<br />

should be compared with Table 5.1.


Conceptual and Ma<strong>the</strong>matical Modell<strong>in</strong>g <strong>in</strong> <strong>River</strong> <strong>Restoration</strong>: Do We Have Unreasonable Confi dence? 73<br />

Table 5.3 Size of 90% confi dence <strong>in</strong>terval (<strong>in</strong> GL/year) for volume of discharge required to mimic natural frequency and duration<br />

of bed scour<strong>in</strong>g events for different sources of error (best estimate of volume is 157 GL/year)<br />

Source of Error Channel Hydraulics Critical Shear Flow Regime<br />

Measurement error neglected 37 20<br />

Model error 234 167 13<br />

Sample error 164 104 107<br />

Comb<strong>in</strong>ed 267 209 125<br />

All sources comb<strong>in</strong>ed 350<br />

is 260 m 3 /s, <strong>the</strong>n to generate <strong>the</strong> natural duration and frequency<br />

of <strong>the</strong>se fl ow pulses would require about 160 GL/<br />

yr, but it could range between 10 GL/year and 360 GL/yr<br />

(i.e. <strong>the</strong> 90th percentile confi dence <strong>in</strong>terval).<br />

The numbers <strong>in</strong> Table 5.3 are <strong>the</strong> confi dence <strong>in</strong>tervals on<br />

this volume of <strong>the</strong> fl ush<strong>in</strong>g fl ow. For example, consider<strong>in</strong>g<br />

all sources of error, <strong>the</strong> 90% confi dence limits are 10 GL/<br />

year and 360 GL/year so <strong>the</strong> magnitude of <strong>the</strong> <strong>in</strong>terval is<br />

350 GL/year which is close to 10% of <strong>the</strong> mean annual<br />

fl ow! Expressed ano<strong>the</strong>r way, 360 GL have to be allocated<br />

for <strong>the</strong> fl ow component <strong>in</strong> order to be 95% confi dent that<br />

<strong>the</strong>re is suffi cient water that <strong>the</strong> bed will be turned over at<br />

a natural frequency and duration, given all of <strong>the</strong> sources of<br />

uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> method. How much is this water worth?<br />

One way to estimate this is to consider <strong>the</strong> cost of <strong>in</strong>frastructure<br />

projects that would be required to achieve comparable<br />

water sav<strong>in</strong>gs (pipel<strong>in</strong>es etc.). The Snowy Water<br />

Inquiry (1998a,b) estimated that environmental fl ows,<br />

costed <strong>in</strong> this way, were worth $US 0.45 million per GL/<br />

year. Clearly this uncerta<strong>in</strong>ty has economic signifi cance.<br />

5.3.4 Discussion<br />

In this simple analysis three o<strong>the</strong>r sources of uncerta<strong>in</strong>ty<br />

have not been considered. The fi rst is <strong>the</strong> simplifi cation<br />

of <strong>the</strong> fl ush<strong>in</strong>g fl ow problem to a s<strong>in</strong>gle threshold shear<br />

stress. The range of τ* used accounts for many of <strong>the</strong><br />

well-known problems of hid<strong>in</strong>g, pack<strong>in</strong>g and o<strong>the</strong>r particle<br />

<strong>in</strong>teractions that affect bed material transport, but <strong>the</strong>re is<br />

still <strong>the</strong> issue of hysteresis. The same discharge on <strong>the</strong><br />

ris<strong>in</strong>g and fall<strong>in</strong>g limb of <strong>the</strong> hydrograph has very different<br />

transport<strong>in</strong>g capacity (Gomez, 1991). It is also possible<br />

that a fl ow exceed<strong>in</strong>g <strong>the</strong> threshold for <strong>in</strong>cipient motion<br />

will be required to mobilize bulk quantities of <strong>the</strong> bed<br />

sediments.<br />

The second uncerta<strong>in</strong>ty not considered is <strong>the</strong> longitud<strong>in</strong>al<br />

variation <strong>in</strong> <strong>the</strong> effect of <strong>the</strong> fl ush<strong>in</strong>g fl ow. In this case<br />

a threshold fl ow <strong>in</strong> a s<strong>in</strong>gle reach has been estimated. The<br />

target fl ow may well turn <strong>the</strong> bed over at <strong>the</strong> sample site,<br />

but what will <strong>the</strong> fl ow do up and downstream of that site?<br />

Given storage effects, a larger fl ow might have to be<br />

released to turn <strong>the</strong> bed over fur<strong>the</strong>r downstream. The<br />

result might be that <strong>the</strong> upstream site is not just ‘turnedover’<br />

but it is progressively scoured away. This <strong>in</strong>troduces<br />

<strong>the</strong> third uncerta<strong>in</strong>ty which is <strong>the</strong> susta<strong>in</strong>ability of <strong>the</strong><br />

process target.<br />

The fl ush<strong>in</strong>g fl ow may work <strong>the</strong> fi rst time, but <strong>the</strong> bed<br />

will <strong>the</strong>n presumably progressively armour, as <strong>the</strong>re may<br />

be no source of coarse sediment below <strong>the</strong> dam. Ei<strong>the</strong>r <strong>the</strong><br />

fl ush<strong>in</strong>g fl ow will have to be progressively <strong>in</strong>creased over<br />

time, or <strong>the</strong> idea of a fl ush<strong>in</strong>g fl ow is simply not susta<strong>in</strong>able<br />

<strong>in</strong> this situation. Such analyses are completed as<br />

steady-state models when <strong>in</strong> reality <strong>the</strong> release of environmental<br />

fl ows will lead to changes <strong>in</strong> <strong>the</strong> <strong>in</strong>put variables.<br />

An example of this problem is <strong>the</strong> major fl ush<strong>in</strong>g-fl ow<br />

experiment carried out on <strong>the</strong> Colorado <strong>River</strong> (Collier<br />

et al., 1997). The fl ood was successful <strong>in</strong> scour<strong>in</strong>g and<br />

rejuvenat<strong>in</strong>g <strong>the</strong> po<strong>in</strong>t-bars of <strong>the</strong> river below <strong>the</strong> Glen<br />

Canyon Dam but this does not mean that it can be successful<br />

<strong>in</strong>defi nitely.<br />

Major sources of uncerta<strong>in</strong>ty have been considered <strong>in</strong> a<br />

reasonably simple geomorphic problem: a fl ush<strong>in</strong>g fl ow.<br />

Most discussion of uncerta<strong>in</strong>ty <strong>in</strong> this type of analysis is<br />

associated with <strong>the</strong> entra<strong>in</strong>ment function, or <strong>in</strong> <strong>the</strong> measurement<br />

of <strong>the</strong> particle size (Kondolf and Wilcock, 1996).<br />

It is <strong>in</strong>terest<strong>in</strong>g to note <strong>the</strong> many sources of uncerta<strong>in</strong>ty<br />

(and error) <strong>in</strong> such analyses. In fact, for this example, <strong>the</strong><br />

ma<strong>in</strong> uncerta<strong>in</strong>ty comes from <strong>the</strong> hydraulics, particularly<br />

<strong>the</strong> estimation of roughness.<br />

It is pert<strong>in</strong>ent to ask what options are available to reduce<br />

<strong>the</strong> uncerta<strong>in</strong>ty? For a geomorphologist <strong>the</strong> response may<br />

be to argue for <strong>the</strong> use of a more sophisticated bed-load<br />

threshold function, but this will not reduce <strong>the</strong> uncerta<strong>in</strong>ty<br />

associated with <strong>the</strong> hydraulics and hydrology. The geomorphologist<br />

could also argue that more research would<br />

reduce uncerta<strong>in</strong>ty. This may be true, but it is also fair to<br />

say that 150 years of research <strong>in</strong>to bed-load transport<br />

re<strong>in</strong>forces <strong>the</strong> view that <strong>the</strong>re is unlikely to be a simple<br />

pr<strong>in</strong>ciple that will dramatically change our capacity to<br />

predict this process. Instead <strong>the</strong> best option is to make<br />

extra measurements at <strong>the</strong> site. The geomorphologists<br />

could ei<strong>the</strong>r measure variables that reduce <strong>the</strong> range of


74 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

uncerta<strong>in</strong>ty <strong>in</strong> each of <strong>the</strong> modelled variables discussed<br />

above, or <strong>the</strong>y could directly measure bed-load movement<br />

to observe <strong>the</strong> discharge at which <strong>the</strong> bed turns over.<br />

Reduc<strong>in</strong>g Model <strong>Uncerta<strong>in</strong>ty</strong><br />

Extra fi eld measurements can reduce <strong>the</strong> uncerta<strong>in</strong>ty and<br />

so <strong>the</strong> amount of water required for <strong>the</strong> fl ush<strong>in</strong>g fl ow. The<br />

options <strong>in</strong>clude:<br />

• extra cross-section surveys;<br />

• record<strong>in</strong>g stages along <strong>the</strong> reach over <strong>the</strong> range of fl ows<br />

be<strong>in</strong>g considered to avoid <strong>the</strong> need to extrapolate<br />

Mann<strong>in</strong>g n from calibration at a s<strong>in</strong>gle discharge; or<br />

• larger bed material samples and more careful measurement<br />

of particle diameter.<br />

Most of <strong>the</strong> error <strong>in</strong> this example lies <strong>in</strong> estimates<br />

of <strong>the</strong> channel hydraulics (Table 5.4). The most effective<br />

way to reduce uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong>se modelled estimates<br />

is to survey water levels over a range of discharges to<br />

allow calibration of Mann<strong>in</strong>g n for a range of fl ows. Some<br />

uncerta<strong>in</strong>ty would rema<strong>in</strong> for extrapolations outside this<br />

range of discharges. Similarly, survey<strong>in</strong>g more cross-sections<br />

will produce a greater decrease <strong>in</strong> error than will a<br />

more accurate estimate of <strong>the</strong> entra<strong>in</strong>ment function.<br />

Overall, <strong>the</strong> estimates are <strong>in</strong>sensitive to uncerta<strong>in</strong>ties <strong>in</strong><br />

<strong>the</strong> fl ow regime, such as an <strong>in</strong>crease <strong>in</strong> <strong>the</strong> length of fl ow<br />

record, or improvements <strong>in</strong> <strong>the</strong> accuracy of <strong>the</strong> rat<strong>in</strong>g<br />

curves.<br />

Ano<strong>the</strong>r way to reduce uncerta<strong>in</strong>ty (and <strong>the</strong> waste of<br />

water) is not to mimic a ‘natural’ fl ow regime. The uncerta<strong>in</strong>ty<br />

(and cost) that this can <strong>in</strong>troduce must be appreciated.<br />

A reductionist approach is to understand <strong>the</strong> target<br />

processes, and release fl ows when required, to achieve this<br />

goal. In <strong>the</strong> case of a fl ush<strong>in</strong>g fl ow, it would be better to<br />

monitor <strong>the</strong> bed processes so that managers knew when<br />

<strong>the</strong> bed needed to be fl ushed (say after a particular series<br />

of lower fl ows). Thus, <strong>the</strong> cost of lost fl ow would be<br />

replaced by <strong>the</strong> cost of monitor<strong>in</strong>g <strong>the</strong> bed.<br />

A clear conclusion from this analysis is that a few extra<br />

measurements can dramatically reduce uncerta<strong>in</strong>ty. In this<br />

particular case, it may be cheaper and more accurate to<br />

simply observe <strong>the</strong> processes directly, ra<strong>the</strong>r than rely on<br />

modell<strong>in</strong>g. The most effi cient way to achieve this is by<br />

trial releases of water from dams (<strong>in</strong> <strong>the</strong> fl ush<strong>in</strong>g-fl ow<br />

case). Dam managers may argue that <strong>the</strong> water is too<br />

valuable to ‘waste’ on such exercises. However, at least<br />

<strong>in</strong> our example, far more water (and hence money) could<br />

be wasted as a consequence of <strong>the</strong> uncerta<strong>in</strong>ty.<br />

However, river restoration projects are often carried out<br />

under tight time constra<strong>in</strong>ts for political or economic<br />

reasons. There is rarely time to monitor <strong>the</strong> processes that<br />

underp<strong>in</strong> <strong>the</strong> modell<strong>in</strong>g or to check <strong>the</strong> predictions. In a<br />

typical stream rehabilitation project, a consultant is<br />

engaged and given perhaps a month or two to report. Any<br />

fi eld <strong>in</strong>vestigations are expected to be brief <strong>in</strong>spections to<br />

develop a conceptual model.<br />

5.4 CONCLUDING DISCUSSION<br />

A central contribution of geomorphology to <strong>the</strong> new practice<br />

of stream restoration is <strong>in</strong> develop<strong>in</strong>g conceptual<br />

models that describe <strong>the</strong> change of stream form and<br />

process over time. This exercise also provides a ‘reference’<br />

condition that can be <strong>the</strong> target for restoration<br />

actions. If <strong>the</strong> conceptual model is wrong, or is too simplistic,<br />

<strong>the</strong>n it does not matter how well executed <strong>the</strong><br />

management actions are, <strong>the</strong> project outcomes rema<strong>in</strong><br />

highly uncerta<strong>in</strong>.<br />

In this chapter <strong>the</strong> development of <strong>the</strong> conceptual geomorphic<br />

model for <strong>the</strong> restoration of <strong>the</strong> lower Snowy<br />

<strong>River</strong> has been described. This is an <strong>in</strong>terest<strong>in</strong>g example<br />

because it illustrates <strong>the</strong> tension between <strong>the</strong> typically<br />

iterative process of geomorphic discovery and <strong>the</strong> need<br />

for action by managers. It also shows <strong>the</strong> dangers of too<br />

much confi dence <strong>in</strong> early <strong>in</strong>terpretations, when <strong>the</strong>se can<br />

trigger expensive management actions that are diffi cult to<br />

reverse even with adaptive management (Stankey et al.,<br />

2003).<br />

<strong>Restoration</strong> can be a very expensive exercise but <strong>the</strong>re<br />

is often little basis for managers to assess how much confi<br />

dence <strong>the</strong>y should have <strong>in</strong> conceptual models that are<br />

presented to <strong>the</strong>m. Five methods have been proposed that<br />

Table 5.4 Size of 90% confi dence <strong>in</strong>terval (expressed <strong>in</strong> GL/year) for volume of discharge required to mimic natural frequency<br />

and duration of bed scour<strong>in</strong>g events with one source of uncerta<strong>in</strong>ty removed (e.g. if completely certa<strong>in</strong> about <strong>the</strong> hydraulic<br />

roughness of <strong>the</strong> channel, <strong>the</strong>n <strong>the</strong> confi dence <strong>in</strong>terval would be reduced from 350 GL/yr down to 280 GL/year, as shown <strong>in</strong> <strong>the</strong><br />

bottom left cell of <strong>the</strong> table)<br />

Source of Error Omitted from Analysis Channel Hydraulics Critical Shear Flow Regime<br />

Sample error Cross-sections 280 No. of particles 330 Years of record 340<br />

Measurement error – Gra<strong>in</strong> size 350 Rat<strong>in</strong>g curve 350<br />

Model error Mann<strong>in</strong>g n 270 Shields entra<strong>in</strong>ment 320 Extrapolat<strong>in</strong>g from gauged<br />

catchments 350


Conceptual and Ma<strong>the</strong>matical Modell<strong>in</strong>g <strong>in</strong> <strong>River</strong> <strong>Restoration</strong>: Do We Have Unreasonable Confi dence? 75<br />

can be used to test <strong>the</strong> validity of conceptual models that<br />

often fall outside <strong>the</strong> area of peer review: <strong>in</strong>dependent<br />

review of <strong>the</strong> model; establish<strong>in</strong>g simple guidel<strong>in</strong>es for<br />

evaluat<strong>in</strong>g <strong>the</strong> uncerta<strong>in</strong>ty of a conceptual model based on<br />

<strong>the</strong> type of strength of evidence provided; ensur<strong>in</strong>g that<br />

an appropriate percentage of <strong>the</strong> total cost of <strong>the</strong> project<br />

is committed to <strong>the</strong> development of <strong>the</strong> conceptual model;<br />

hold<strong>in</strong>g back some of <strong>the</strong> <strong>in</strong>formation and data that underp<strong>in</strong><br />

<strong>the</strong> model, and us<strong>in</strong>g <strong>the</strong>se data to verify <strong>the</strong> hypo<strong>the</strong>tical<br />

model; and pass<strong>in</strong>g all of <strong>the</strong> data and <strong>in</strong>formation to<br />

a third party, without <strong>in</strong>terpretation, so that a compet<strong>in</strong>g<br />

model can be developed.<br />

The feasibility has been demonstrated of quantify<strong>in</strong>g<br />

uncerta<strong>in</strong>ty <strong>in</strong> a geomorphic design model, us<strong>in</strong>g <strong>the</strong> case<br />

study of a fl ow to fl ush <strong>the</strong> fi ne sediments from <strong>the</strong> bed of<br />

<strong>the</strong> Goulburn <strong>River</strong>. In this case, <strong>the</strong> uncerta<strong>in</strong>ty was very<br />

large. Surpris<strong>in</strong>gly, <strong>the</strong> ma<strong>in</strong> source of uncerta<strong>in</strong>ty came<br />

from <strong>the</strong> hydraulic modell<strong>in</strong>g and relatively little from<br />

ei<strong>the</strong>r <strong>the</strong> bed-load entra<strong>in</strong>ment function or <strong>the</strong> hydrological<br />

modell<strong>in</strong>g of <strong>the</strong> fl ow series. It is also strik<strong>in</strong>g how<br />

add<strong>in</strong>g extra fi eld measurements can reduce <strong>the</strong> uncerta<strong>in</strong>ty<br />

<strong>in</strong> <strong>the</strong> model estimates, particularly <strong>in</strong> estimat<strong>in</strong>g<br />

Mann<strong>in</strong>g n. In fact, <strong>the</strong> key to uncerta<strong>in</strong>ty often comes<br />

down to modell<strong>in</strong>g versus monitor<strong>in</strong>g.<br />

It is necessary to <strong>in</strong>clude some form of data ga<strong>the</strong>r<strong>in</strong>g<br />

with any geomorphic modell<strong>in</strong>g exercises, <strong>in</strong>clud<strong>in</strong>g<br />

modell<strong>in</strong>g geomorphic response to river restoration. It has<br />

been shown how an uncerta<strong>in</strong>ty analysis might be used to<br />

direct this data ga<strong>the</strong>r<strong>in</strong>g effort to most effectively reduce<br />

uncerta<strong>in</strong>ty <strong>in</strong> model predictions. It seems logical that<br />

modell<strong>in</strong>g and data ga<strong>the</strong>r<strong>in</strong>g should be <strong>in</strong>tegrated to m<strong>in</strong>imise<br />

uncerta<strong>in</strong>ties <strong>in</strong> restoration. In many cases, some<br />

form of monitor<strong>in</strong>g is carried out as part of <strong>the</strong> restoration<br />

implementation to check that <strong>the</strong> project achieves what<br />

was <strong>in</strong>tended. However, <strong>the</strong>se monitor<strong>in</strong>g programs are<br />

rarely designed to verify or improve <strong>the</strong> models used <strong>in</strong><br />

plann<strong>in</strong>g <strong>the</strong> restoration. <strong>Uncerta<strong>in</strong>ty</strong> analysis appears to<br />

be useful for optimis<strong>in</strong>g such monitor<strong>in</strong>g programs to<br />

provide improved models for subsequent restoration<br />

decisions.<br />

In some cases <strong>the</strong> most effective way to reduce uncerta<strong>in</strong>ties<br />

is to run trial programs <strong>in</strong> <strong>the</strong> actual systems be<strong>in</strong>g<br />

restored. Even <strong>in</strong> <strong>the</strong>se cases, <strong>the</strong> model and uncerta<strong>in</strong>ty<br />

analysis can provide <strong>the</strong> basis for design<strong>in</strong>g <strong>the</strong> trial to<br />

ensure that it tackles <strong>the</strong> key sources of uncerta<strong>in</strong>ty <strong>in</strong><br />

model predictions. In <strong>the</strong> case of <strong>the</strong> lower Snowy <strong>River</strong>,<br />

<strong>the</strong> state government has recognised <strong>the</strong> value of additional<br />

geomorphic measurements and analysis, and<br />

<strong>in</strong>vested <strong>in</strong> a restoration trial which <strong>in</strong>cludes fi eld measurements<br />

and physical modell<strong>in</strong>g to improve <strong>the</strong> geomorphic<br />

design. It is possible to design monitor<strong>in</strong>g activities<br />

without regard to uncerta<strong>in</strong>ties dur<strong>in</strong>g <strong>the</strong> plann<strong>in</strong>g phase<br />

of a restoration project. However, it is proposed that<br />

stronger <strong>in</strong>tegration of monitor<strong>in</strong>g and modell<strong>in</strong>g activities<br />

will lead to greater improvements <strong>in</strong> <strong>the</strong> knowledge<br />

underp<strong>in</strong>n<strong>in</strong>g river restoration.<br />

Geomorphic models, like models developed <strong>in</strong> any<br />

science, are subject to some uncerta<strong>in</strong>ty. Although this<br />

uncerta<strong>in</strong>ty is widely acknowledged it is rarely evaluated.<br />

We conclude that <strong>the</strong>re is unreasonable confi dence <strong>in</strong><br />

geomorphic models used for river restoration. Certa<strong>in</strong>ly<br />

we were surprised by <strong>the</strong> magnitude of uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong><br />

fl ush<strong>in</strong>g fl ow example and we have been <strong>in</strong>volved <strong>in</strong> many<br />

of <strong>the</strong>se modell<strong>in</strong>g studies <strong>in</strong> Australia. There is rarely any<br />

thought given to <strong>the</strong> best approach to modell<strong>in</strong>g, <strong>in</strong>clud<strong>in</strong>g<br />

<strong>the</strong> calculation of <strong>in</strong>put parameters to m<strong>in</strong>imise uncerta<strong>in</strong>ties<br />

<strong>in</strong> restoration decisions. There is a need to reth<strong>in</strong>k our<br />

approach to geomorphic modell<strong>in</strong>g <strong>in</strong> <strong>the</strong> context of river<br />

restoration, <strong>in</strong> particular <strong>the</strong> benefi ts of evaluat<strong>in</strong>g uncerta<strong>in</strong>ties<br />

<strong>in</strong> both our conceptual and ma<strong>the</strong>matical models.<br />

This requires more expertise, <strong>in</strong>formation and fund<strong>in</strong>g<br />

but <strong>the</strong> result will be more realistic expectations by those<br />

<strong>in</strong>volved <strong>in</strong> <strong>the</strong> restoration project and a more careful<br />

approach to optimis<strong>in</strong>g data ga<strong>the</strong>r<strong>in</strong>g for calculat<strong>in</strong>g<br />

model parameters and verify<strong>in</strong>g model structure.<br />

REFERENCES<br />

Beven K. 2001. Ra<strong>in</strong>fall–Runoff Modell<strong>in</strong>g: The Primer. John<br />

Wiley & Sons Ltd: Chichester.<br />

Brizga SO, F<strong>in</strong>layson BL. 1992. The Snowy <strong>River</strong> Sediment<br />

Study: Investigation <strong>in</strong>to <strong>the</strong> Distribution, Transport and<br />

Sources of Sand <strong>in</strong> <strong>the</strong> Snowy <strong>River</strong> between Lake J<strong>in</strong>dabyne<br />

and Jarrahmond. Department of Water Resources: Melbourne,<br />

Victoria, Australia.<br />

Brizga SO, F<strong>in</strong>layson BL. 1994. Interactions between upland<br />

catchment and lowland rivers: an applied Australian case<br />

study. Geomorphology 9: 189–201.<br />

Brookes A. 1995. Challenges and objectives for geomorphology<br />

<strong>in</strong> UK river management. Earth Surface Processes and Landforms<br />

20: 593–610.<br />

Brookes A, Sear DA. 1996. Geomorphological pr<strong>in</strong>ciples for<br />

restor<strong>in</strong>g channels. In: Brookes A, Shields FD (Eds), <strong>River</strong><br />

Channel <strong>Restoration</strong>: Guid<strong>in</strong>g Pr<strong>in</strong>ciples for Susta<strong>in</strong>able<br />

Projects. John Wiley & Sons Ltd: Chichester; 75–101.<br />

Brooks AP, Brierley GJ, Millar RG. 2003. The long-term control<br />

of vegetation and woody debris on channel and fl oodpla<strong>in</strong><br />

evolution: <strong>in</strong>sights from a paired catchment study <strong>in</strong> sou<strong>the</strong>astern<br />

Australia. Geomorphology 51: 7–29.<br />

Brooks AP, Brierley GJ. 1997. Geomorphic responses of Lower<br />

Bega <strong>River</strong> to catchment disturbance, 1851–1926. Geomorphology<br />

18: 291–304.<br />

Buffi ngton JM, Montgomery DR. 1997. A systematic analysis of<br />

eight decades of <strong>in</strong>cipient motion studies, with special reference<br />

to gravel-bedded rivers. Water Resources Research 33<br />

(8): 1993–2029.<br />

Clarke RT. 1999. <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> <strong>the</strong> estimation of mean annual<br />

fl ood due to rat<strong>in</strong>g curve <strong>in</strong>defi nition. Journal of Hydrology<br />

222: 185–190.


76 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Collier MP, Webb RH, Andrews ED. 1997. Experimental fl ood<strong>in</strong>g<br />

<strong>in</strong> Grand Canyon. Scientifi c American 276: 82–89.<br />

Cooke RV, Reeves RW. 1976. Arroyos and Environmental<br />

Change <strong>in</strong> <strong>the</strong> American Southwest. Oxford Research Studies<br />

<strong>in</strong> Geography, Oxford.<br />

Cullen P. 1989. The Turbulent Boundary Between Science and<br />

Management. Canberra College of Advanced Education:<br />

Canberra, Australia.<br />

DNRE 2002. The FLOWS method: A method for determ<strong>in</strong><strong>in</strong>g<br />

environmental water requirements <strong>in</strong> Victoria Victorian<br />

Department of Natural Resources and Environment,<br />

Australia.<br />

Ersk<strong>in</strong>e WD, Terrazzolo N, Warner RF. 1999. <strong>River</strong> restoration<br />

from <strong>the</strong> hydrogeomorphic impacts of a large hydro-electric<br />

power project: Snowy <strong>River</strong>, Australia. Regulated <strong>River</strong>s 15:<br />

3–24.<br />

Ersk<strong>in</strong>e WD, Tilleard JW. 1997. Formative processes of alternat<strong>in</strong>g,<br />

bank-attached side bars and associated pool-riffl e<br />

sequences on sand-bed streams similar to <strong>the</strong> Snowy <strong>River</strong> at<br />

Jarrahmond, Victoria. Department of Natural Resources and<br />

Environment: Melbourne, Victoria, Australia.<br />

F<strong>in</strong>layson BL, Bird JF. 1989. Initial <strong>in</strong>vestigation <strong>in</strong>to <strong>the</strong> extent<br />

and nature of <strong>the</strong> current sedimentation problem on <strong>the</strong> lower<br />

Snowy <strong>River</strong>. Centre for Environmental Applied Hydrology,<br />

The University of Melbourne: Melbourne, Australia.<br />

Gilbert GK. 1917. Hydraulic m<strong>in</strong><strong>in</strong>g debris <strong>in</strong> <strong>the</strong> Sierra Nevada.<br />

United States Geological Survey Professional Paper 105:<br />

1–154.<br />

Gilvear DJ. 1999. Fluvial geomorphology and river eng<strong>in</strong>eer<strong>in</strong>g:<br />

future roles utiliz<strong>in</strong>g a fl uvial hydrosystems framework. Geomorphology<br />

31: 229–245.<br />

Gippel CJ. 2002. The Victorian Snowy <strong>River</strong>: Review of Historical<br />

Environmental Change and Proposed <strong>Restoration</strong> Options.<br />

East Gippsland Catchment Management Authority: Victoria,<br />

Australia.<br />

Gippel CJ, Anderson BA, Marsh N. 2002. Trial of Snowy <strong>River</strong><br />

<strong>Restoration</strong> Concept Plan: Scop<strong>in</strong>g Study to Review Struc -<br />

tural <strong>Restoration</strong> Options. Report by Fluvial Systems<br />

Pty Ltd, Stockton, to Snowy <strong>River</strong> <strong>Restoration</strong> Project, East<br />

Gippsland Catchment Management Authority: Victoria,<br />

Australia.<br />

Gomez B. 1991. Bedload transport. Earth-Science Reviews 31:<br />

89–132.<br />

Hicks DM, Mason PD. 1998. Roughness characteristics of New<br />

Zealand <strong>River</strong>s, National Institute of Water and Atmospheric<br />

Research, Water Resource Publications: Englewood, New<br />

Zealand.<br />

ID&A 1998. <strong>River</strong> restoration concept plan for <strong>the</strong> Snowy <strong>River</strong><br />

<strong>in</strong> Victoria. Wangaratta, Victoria, Consultants report to East<br />

Gippsland Catchment Management Authority: Victoria,<br />

Australia.<br />

Kondolf GM. 2000. Some suggested guidel<strong>in</strong>es for geomorphic<br />

aspects of anadromous salmonid habitat restoration proposals.<br />

<strong>Restoration</strong> Ecology 8 (1): 48–56.<br />

Kondolf GM, Micheli ER. 1995. Evaluat<strong>in</strong>g stream restoration<br />

projects. Environmental Management 19 (1): 1–15.<br />

Kondolf GM, Smeltzer MW, Railsback SF. 2001. Design and<br />

performance of a channel reconstruction project <strong>in</strong> a coastal<br />

California gravel-bed stream. Environmental Management 28<br />

(6): 761–776.<br />

Kondolf GM, Wilcock PR. 1996. The fl ush<strong>in</strong>g fl ow problem:<br />

defi n<strong>in</strong>g and evaluat<strong>in</strong>g objectives. Water Resources Research<br />

32 (8): 2589–2599.<br />

Kuhnle RA, Alonso CV, Shields FD. 2002. Local scour associated<br />

with angled spur dikes. Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g<br />

128 (12): 1087–1093.<br />

Ladson AR, Argent RM. 2002. Adaptive management of environmental<br />

fl ows: lessons for <strong>the</strong> Murray-Darl<strong>in</strong>g Bas<strong>in</strong> from<br />

three large North American <strong>River</strong>s. Australian Journal of<br />

Water Resources 5 (1): 89–102<br />

Larson M, Goldsmith W. 1997. Incised channel stabilization and<br />

enhancement <strong>in</strong>tegrat<strong>in</strong>g geomorphology and bioeng<strong>in</strong>eer<strong>in</strong>g.<br />

In: Wang SSY, Langendoen EJ, Shields FD (Eds), Management<br />

of Landscapes Disturbed by Channel Incision, University<br />

of Mississippi: Oxford, Mississippi; 458–465.<br />

Manly BFJ. 1997. Randomization, Bootstrap and Monte Carlo<br />

Methods <strong>in</strong> Biology. Chapman and Hall: London.<br />

Milhous RT. 1995. Flush<strong>in</strong>g fl ows for habitat restoration. In:<br />

Espey W (Ed) Water Resources Eng<strong>in</strong>eer<strong>in</strong>g, American<br />

Society of Civil Eng<strong>in</strong>eers: San Antonio, Texas; 663–667.<br />

Miller JR, Ritter JB. 1996. An exam<strong>in</strong>ation of <strong>the</strong> Rosgen classifi<br />

cation of natural rivers. Catena 27: 295–299.<br />

Neilsen MB. 1996. Lowland stream restoration <strong>in</strong> Denmark. In:<br />

Brookes A, Shields FD (Eds), <strong>River</strong> Channel <strong>Restoration</strong>:<br />

Guid<strong>in</strong>g Pr<strong>in</strong>ciples for Susta<strong>in</strong>able Projects. John Wiley &<br />

Sons Ltd: Chichester; 269–290.<br />

Newbury R, Gaboury M. 1993. Stream analysis and fi sh habitat<br />

design – a fi eld manual. Newbury Hydraulics Ltd: British<br />

Columbia, Canada.<br />

O’Neill MP, Kuhns MR. 1994. Stream bank erosion and fl ush<strong>in</strong>g<br />

fl ows. Stream Notes (July), USDA Forest Service.<br />

Owen R. 1997. The Lower Snowy <strong>River</strong> Revegetation Strategy.<br />

Snowy <strong>River</strong> Improvement Trust: Nungurner Hills Nursery,<br />

Victoria, Australia.<br />

Poff NL, Allan JD, Ba<strong>in</strong> MB et al. 1997. The natural fl ow regime,<br />

a paradigm for river conservation and restoration. BioScience<br />

47: 769–84.<br />

Raadick TA, O’Connor JP. 1997. Fish and Decapod Crustacean<br />

Survey, and Habitat Assessment, of <strong>the</strong> Lower Snowy <strong>River</strong>,<br />

Victoria. Report to <strong>the</strong> Snowy <strong>River</strong> Improvement Trust by <strong>the</strong><br />

Freshwater Ecology Division of <strong>the</strong> Department of Natural<br />

Resources.<br />

Regan HM, Colyvan M, Burgman M. 2002. A taxonomy of<br />

uncerta<strong>in</strong>ty for ecology and conservation biology. Ecological<br />

Applications 12 (2): 618–628.<br />

Richter BD, Baumgartner JV, Wig<strong>in</strong>gton R, Braun DP. 1997.<br />

How much water does a river need? Freshwater Biology 37:<br />

231–249.<br />

Rosgen DL. 1996. Applied <strong>River</strong> Morphology. Wildland Hydrology:<br />

Pagosa Spr<strong>in</strong>gs, Colorado.<br />

Ru<strong>the</strong>rfurd ID. 2001. Storage and movement of slugs of sand <strong>in</strong><br />

a large catchment: develop<strong>in</strong>g a plan to rehabilitate <strong>the</strong> Glenelg


Conceptual and Ma<strong>the</strong>matical Modell<strong>in</strong>g <strong>in</strong> <strong>River</strong> <strong>Restoration</strong>: Do We Have Unreasonable Confi dence? 77<br />

<strong>River</strong>, SE Australia. In: Anthony DJ, Harvey MD, Laronne JB,<br />

Mosley MP (Eds), Apply<strong>in</strong>g Geomorphology to Environmental<br />

Management. Water Resources Publications: Denver, Colorado;<br />

309–332.<br />

Schumm SA, Harvey MD, Watson CC. 1984. Incised Channels:<br />

Morphology, Dynamics and Control. Water Resources Publications:<br />

Littleton, Colorado.<br />

Sear DA. 1993. F<strong>in</strong>e sediment <strong>in</strong>fi ltration <strong>in</strong>to gravel spawn<strong>in</strong>g<br />

beds with<strong>in</strong> a regulated river experienc<strong>in</strong>g fl oods: Ecological<br />

implications for salmonids. Regulated <strong>River</strong>s 8: 373–390.<br />

Sear DA. 1994. <strong>River</strong> restoration and geomorphology. Aquatic<br />

Conservation: Mar<strong>in</strong>e and Freshwater Ecosystems 4:<br />

169–177.<br />

Seddon GS. 1994. Search<strong>in</strong>g for <strong>the</strong> Snowy, an Environmental<br />

History. Allen and Unw<strong>in</strong>: St Leonards, New South Wales.<br />

Simon A. 1989. A model of channel response <strong>in</strong> disturbed<br />

alluvial channels. Earth Surface Processes and Landforms 14:<br />

11–26.<br />

Snowy Water Inquiry 1998a. Snowy Water Inquiry: Draft Options<br />

for Discussion. Snowy Water Inquiry: Sydney, NSW,<br />

Australia.<br />

Snowy Water Inquiry1998b. Appendix of Resource Materials<br />

(Part 2). Snowy Water Inquiry: Sydney, NSW, Australia.<br />

Stankey GH, Bormann, BT, Ryan C. 2003. Adaptive management<br />

and <strong>the</strong> northwest forest plan: Rhetoric and reality.<br />

Journal of Forestry 40: 40–46.<br />

Stewardson MJ. 1998. Pool formation – fl uvial processes, Section<br />

6 <strong>in</strong> ID&A, <strong>River</strong> restoration concept plan for <strong>the</strong> Snowy <strong>River</strong><br />

<strong>in</strong> Victoria. Report to East Gippsland Catchment Management<br />

Authority by ID&A: Wangaratta, Victoria, Australia.<br />

Stewardson MJ, Anderson B. 2002. Variations <strong>in</strong> <strong>the</strong> fl ow resistance<br />

of natural channels with discharge. Proceed<strong>in</strong>gs of <strong>the</strong><br />

Hydrology and Water Resources Symposium. Institution of<br />

Eng<strong>in</strong>eers: Melbourne, Australia.<br />

Stewardson MJ, Cott<strong>in</strong>gham P. 2002. A demonstration of <strong>the</strong><br />

fl ow events method: Environmental fl ow requirements of <strong>the</strong><br />

Broken <strong>River</strong>. Australian Journal of Water Resources 5 (1):<br />

35–48.<br />

Strom HG. 1936. The Flood Problem <strong>in</strong> Gippsland, 2. The Snowy<br />

<strong>River</strong>. State <strong>River</strong>s and Water Supply Commission: Melbourne,<br />

Victoria, Australia.<br />

Volkman JM. 1999. How do you learn from a river? <strong>Manag<strong>in</strong>g</strong><br />

uncerta<strong>in</strong>ty <strong>in</strong> species conservation policy. Wash<strong>in</strong>gton Law<br />

Review 74: 719–762.<br />

Walters C. 1997. Challeges <strong>in</strong> adaptive management of riparian<br />

and coastal ecosystems. Conservation Ecology [onl<strong>in</strong>e] URL<br />

http://www.consecol.org/vol1/iss2/art1.<br />

APPENDIX 5.1<br />

Flows at <strong>the</strong> Goulburn <strong>River</strong> site are regulated by operation<br />

of Lake Eildon, a large water supply reservoir, upstream of<br />

<strong>the</strong> study site. Flow data for gaug<strong>in</strong>g stations located on<br />

n<strong>in</strong>e unregulated tributaries upstream of <strong>the</strong> Goulburn<br />

<strong>River</strong> site were used to model natural daily fl ows at this site<br />

(i.e. fl ows that would occur <strong>in</strong> <strong>the</strong> absence of Lake Eildon).<br />

Seven of <strong>the</strong>se tributaries are upstream of Lake Eildon and<br />

two o<strong>the</strong>r tributaries have confl uences with <strong>the</strong> Goulburn<br />

<strong>River</strong> between Lake Eildon and <strong>the</strong> survey site (Table 5.5).<br />

In total 80% of <strong>the</strong> 4225 km2 catchment at <strong>the</strong> survey site<br />

is upstream of <strong>the</strong>se fl ow gauges (510 km2 upstream Eildon<br />

and 314 km2 between Eildon and <strong>the</strong> study site). Flows <strong>in</strong><br />

<strong>the</strong> ungauged portion of <strong>the</strong> catchments upstream and<br />

downstream of Lake Eildon are estimated by scal<strong>in</strong>g fl ow<br />

<strong>in</strong> <strong>the</strong> gauged portion of <strong>the</strong> catchments upstream and<br />

downstream of Lake Eildon respectively. The scal<strong>in</strong>g<br />

factor was estimated us<strong>in</strong>g a l<strong>in</strong>ear function of <strong>the</strong> ratio of<br />

gauged and ungauged catchment areas, derived from <strong>the</strong><br />

available streamfl ow data (Figure 5.5). Daily fl ows at<br />

<strong>the</strong> survey site are estimated as <strong>the</strong> sum of daily fl ows (for<br />

<strong>the</strong> same day) at <strong>the</strong> upstream gauges and estimated <strong>in</strong> <strong>the</strong><br />

ungauged catchments. Rout<strong>in</strong>g effects (i.e. travel times<br />

and attenuation of fl ood peaks) are neglected because <strong>the</strong>re<br />

Table 5.5 Gauged tributaries of <strong>the</strong> Goulburn <strong>River</strong> upstream of <strong>the</strong> Goulburn <strong>River</strong><br />

survey site<br />

Tributary Catchment area (km 2 )<br />

Upstream of Lake Eildon<br />

Delatite <strong>River</strong> 368<br />

Howqua <strong>River</strong> 368<br />

Jamison <strong>River</strong> 368<br />

Goulburn <strong>River</strong> at Dohertys Infl ow 694<br />

Big <strong>River</strong> 619<br />

Ford Creek 115<br />

Brankeet Creek 121<br />

Between Lake Eildon and <strong>the</strong> surveys site<br />

Rubicon <strong>River</strong> 129<br />

Acheron <strong>River</strong> 619


78 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

is no <strong>in</strong>formation with which to establish a rout<strong>in</strong>g<br />

model.<br />

The regression equation <strong>in</strong> Figure 5.5 is used to estimate<br />

a scal<strong>in</strong>g factor for estimat<strong>in</strong>g fl ows <strong>in</strong> <strong>the</strong> ungauged catchments.<br />

Data po<strong>in</strong>ts <strong>in</strong> this plot show <strong>the</strong> mean ration of fl ow<br />

Statistic of daily flow ratio<br />

0.5<br />

0.45<br />

0.4<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

y = 1.33x - 0.05<br />

R 2 = 0.87<br />

Table 5.6 Characteristics of <strong>the</strong> ratio of daily fl ows at each of seven streamfl ow gauges to <strong>the</strong> sum of daily fl ows at <strong>the</strong> o<strong>the</strong>r six<br />

gauges (Gauges are all located upstream of Lake Eildon)<br />

Gauge Catchment<br />

Area (km 2 )<br />

y = 0.35x - 0.0037<br />

R 2 = 0.83<br />

0 0.1 0.2 0.3 0.4<br />

Ratio catchment areas<br />

mean of daily flow ratio<br />

standard dev. of daily flow ratio<br />

Figure 5.5 Mean and standard deviation of daily fl ows ratio and<br />

catchment area ratios for seven streamfl ow gauges upstream of<br />

Lake Eildon. Ratios are calculated at each gauge <strong>in</strong> turn by divid<strong>in</strong>g<br />

daily fl ows and catchment area at <strong>the</strong> gauge by <strong>the</strong> sum of<br />

daily fl ows and catchment areas at <strong>the</strong> o<strong>the</strong>r six gauges<br />

Mean Flow<br />

(ML/day)<br />

Catchment<br />

Area Ratio<br />

Mean of<br />

Daily Flow<br />

Ratio<br />

Standard<br />

Deviation of Daily<br />

Flow Ratio<br />

Characteristics for <strong>the</strong> streamfl ow gauges upstream of Lake Eildon<br />

Delatite <strong>River</strong> 368 297 0.16 0.10 0.037 0.66<br />

Howqua <strong>River</strong> 368 477 0.16 0.18 0.051 0.94<br />

Jamison <strong>River</strong> 368 560 0.16 0.21 0.056 0.91<br />

Goulburn <strong>River</strong> 694 882 0.35 0.35 0.097 0.93<br />

Big <strong>River</strong> 619 843 0.30 0.46 0.14 0.93<br />

Ford Creek 115 34 0.045 0.0049 0.011 0.44<br />

Brankeet Creek 121 48 0.048 0.022 0.018 0.56<br />

Characteristics estimated for <strong>the</strong> ungauged catchments us<strong>in</strong>g regression equation <strong>in</strong> Figure 5.5<br />

Upstream of Eildon 510 – 0.19 0.21 0.063 0.9*<br />

Eildon to study site 314 – 0.42 0.51 0.14 0.9*<br />

* selected based on serial correlations derived from streamfl ow gauge data<br />

at one of <strong>the</strong> gauges to sum of fl ow at all o<strong>the</strong>r gauges. In<br />

reality this ratio varies from day to day. The mean, standard<br />

deviation and lag-1 serial correlation for this ratio<br />

were calculated for each of <strong>the</strong> seven sites upstream of<br />

Lake Eildon (Table 5.6). These characteristics were estimated<br />

for <strong>the</strong> ungauged catchments us<strong>in</strong>g <strong>the</strong> regression<br />

equation <strong>in</strong> Figure 5.5. A stochastic model is used to represent<br />

uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> fl ow estimated <strong>in</strong> <strong>the</strong> ungauged<br />

catchment. This model gives fl ow from <strong>the</strong> ungauged<br />

catchment on <strong>the</strong> ith day of <strong>the</strong> 25 year record as:<br />

Q = [ mr . + ( 1−m) r 1] Q and r = N(<br />

ms , ) (5.4)<br />

ui , i i− ∑<br />

j<br />

ji ,<br />

i<br />

where Qj,i is <strong>the</strong> fl ow at <strong>the</strong> jth gauge site on <strong>the</strong> ith day.<br />

Values of m were calibrated as 0.1 to provide a lag-1 serial<br />

correlation of 0.9 for <strong>the</strong> fl ow ratios <strong>in</strong> ungauged catchments<br />

upstream and downstream of Lake Eildon. The<br />

daily parameter r i is a normally distributed random variable.<br />

The mean (µ) of r i was set equal to <strong>the</strong> mean of daily<br />

fl ow ratios estimated from <strong>the</strong> regression <strong>in</strong> Figure 5.5 The<br />

standard deviation (σ) was adjusted to replicate <strong>the</strong> standard<br />

deviations calculated from <strong>the</strong> regressions <strong>in</strong> Figure<br />

5.5. Standard deviations chosen for upstream and downstream<br />

of Lake Eildon were 0.27 and 0.61 respectively.<br />

Note that <strong>the</strong>se are higher than <strong>the</strong> values estimated<br />

directly from <strong>the</strong> regression equations to account for <strong>the</strong><br />

effect of lag correlation <strong>in</strong> <strong>the</strong> stochastic model. This stochastic<br />

model is used to generate replicate 25-year timeseries<br />

of fl ow from <strong>the</strong> ungauged catchments.<br />

Lag-1 Serial Correlation<br />

of Daily Flow Ratio


6.1 INTRODUCTION: THE CASE<br />

FOR RESTORATION<br />

6<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and<br />

Floodpla<strong>in</strong> <strong>Restoration</strong><br />

Franc<strong>in</strong>e M.R. Hughes 1 , Timothy Moss 2 and Keith S. Richards 3<br />

1 Department of Life Sciences, Anglia Rusk<strong>in</strong> University, UK<br />

2 Institute for Regional Development and Structural Plann<strong>in</strong>g (IRS), Germany<br />

3 Department of Geography, University of Cambridge, UK<br />

The riparian and fl oodpla<strong>in</strong> zones of rivers are physically<br />

dynamic places, subject to <strong>the</strong> delivery and removal of<br />

water and sediments dur<strong>in</strong>g fl ood events. Ecosystems that<br />

occupy <strong>the</strong>se places have evolved a tolerance to <strong>the</strong>se<br />

natural disturbance processes and many of <strong>the</strong>ir component<br />

species have become dependent on <strong>the</strong>m for completion<br />

of <strong>the</strong>ir life cycles. In many river valleys, riparian and<br />

fl oodpla<strong>in</strong> zones would once have been occupied by forested<br />

ecosystems, composed of a dynamic mosaic of<br />

forest types <strong>in</strong> different successional stages, <strong>in</strong>terspersed<br />

with more open wetland communities with emergent vegetation.<br />

Such fl oodpla<strong>in</strong> forests have high levels of biodiversity<br />

because <strong>the</strong>y are at <strong>the</strong> <strong>in</strong>terface between terrestrial<br />

and lotic ecosystems (Petts, 1990). In addition, <strong>the</strong>y experience<br />

frequent disturbance from fl oods which create <strong>the</strong><br />

conditions for a heterogeneous mosaic of habitats across<br />

<strong>the</strong> fl oodpla<strong>in</strong>, each support<strong>in</strong>g a varied mix of species<br />

(Nilsson et al., 1991a; Hughes et al., 2005). A high plant<br />

species diversity has been recorded on fl oodpla<strong>in</strong>s from<br />

rivers <strong>in</strong> many different bioclimatic zones. For example,<br />

fl oodpla<strong>in</strong>s <strong>in</strong> <strong>the</strong> Amazon bas<strong>in</strong> account for 20% of tree<br />

species diversity (Junk et al., 1989). In <strong>the</strong> Tana <strong>River</strong><br />

fl oodpla<strong>in</strong> forests <strong>in</strong> Kenya, which stretch for only 200 km<br />

of river length and average only 1 km <strong>in</strong> width, 175 woody<br />

plant species, over 250 species of birds and at least 57<br />

species of mammals have been found, <strong>in</strong>clud<strong>in</strong>g two<br />

endemic primates (Medley and Hughes, 1996).<br />

Riparian and fl oodpla<strong>in</strong> ecosystems not only have<br />

<strong>in</strong>tr<strong>in</strong>sic and <strong>in</strong>tangible values associated with <strong>the</strong>se high<br />

levels of biodiversity, and with <strong>the</strong>ir diverse landscape<br />

character, but <strong>the</strong>y can also be valued because <strong>the</strong>y contribute<br />

to timber production and carbon sequestration,<br />

fl oodwater storage, groundwater recharge, pollution<br />

control and even recreation. The importance of riparian<br />

and fl oodpla<strong>in</strong> zones can thus lie <strong>in</strong> this wide range of<br />

natural functions and services that <strong>the</strong>y provide, although<br />

<strong>the</strong>re is uncerta<strong>in</strong>ty about how <strong>the</strong>se can be valued over<br />

time, given <strong>the</strong>ir dynamic and vary<strong>in</strong>g nature. Such valuation<br />

is today <strong>in</strong>creas<strong>in</strong>gly necessary, s<strong>in</strong>ce hydrological<br />

pathways <strong>in</strong> river bas<strong>in</strong>s have often been altered <strong>in</strong>directly<br />

through land-use change and directly through management<br />

of <strong>the</strong> fl ow regime. In downstream fl oodpla<strong>in</strong> zones<br />

<strong>in</strong> particular <strong>the</strong>re have been many eng<strong>in</strong>eered changes to<br />

river channels, lead<strong>in</strong>g to isolation of fl oodpla<strong>in</strong>s from<br />

<strong>the</strong>ir channels and to severe damage or eradication of<br />

fl oodpla<strong>in</strong> ecosystems. All <strong>the</strong>se changes have severely<br />

limited or even destroyed <strong>the</strong> capacity of fl oodpla<strong>in</strong>s to<br />

deliver <strong>the</strong>ir natural functions and services.<br />

The disappearance of <strong>the</strong>se ecosystems from <strong>the</strong> landscape<br />

is poignantly illustrated by <strong>the</strong> case of fl oodpla<strong>in</strong><br />

forests <strong>in</strong> Europe. 90% of <strong>the</strong>se forested ecosystems have<br />

disappeared and rema<strong>in</strong><strong>in</strong>g patches are often <strong>in</strong> critical<br />

condition. They are listed <strong>in</strong> Annexe I of <strong>the</strong> European<br />

Habitats Directive (92/43/EEC, 1992) as a priority habitat<br />

type and are <strong>in</strong>cluded <strong>in</strong> <strong>the</strong> Natura 2000 network of nature<br />

reserves. In western Europe <strong>the</strong>y are more reduced <strong>in</strong><br />

extent than <strong>in</strong> eastern and central Europe where some<br />

<strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat Edited by Stephen Darby and David Sear<br />

© 2008 John Wiley & Sons, Ltd


80 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Figure 6.1 Map of rema<strong>in</strong><strong>in</strong>g European fl oodpla<strong>in</strong> forests (based on data from UNEP – World Conservation Monitor<strong>in</strong>g Centre <strong>in</strong><br />

UNEP–WCMC, 2000 and Girel et al., 2003)<br />

impressive patches rema<strong>in</strong> (Figure 6.1). Even here, <strong>the</strong>ir<br />

extent is not easy to gauge as many of <strong>the</strong> areas marked<br />

as fl oodpla<strong>in</strong> forest have <strong>in</strong> fact been converted to areas<br />

of forestry on fl oodpla<strong>in</strong>s, without any of <strong>the</strong> characteristic<br />

dynamic features of a naturally function<strong>in</strong>g fl oodpla<strong>in</strong><br />

forest and frequently dom<strong>in</strong>ated by non-native species.<br />

For example, <strong>in</strong> Hungary, where fl ood control works have<br />

reduced <strong>the</strong> fl oodpla<strong>in</strong> area across all river systems from<br />

2.3 billion km 2 to only 1500 km 2 , 40% of <strong>the</strong> rema<strong>in</strong><strong>in</strong>g<br />

areas of fl oodpla<strong>in</strong> forests have been converted to forestry<br />

plantations (Haraszthy, 2001).<br />

6.2 POLICY-RELATED WINDOWS<br />

OF OPPORTUNITY<br />

However, s<strong>in</strong>ce <strong>the</strong> mid-1990s, shifts <strong>in</strong> policy content and<br />

style <strong>in</strong> <strong>the</strong> fi elds of fl ood protection, nature conservation<br />

and agriculture at European Union (EU) and national<br />

levels are creat<strong>in</strong>g ‘w<strong>in</strong>dows of opportunity’ for fl oodpla<strong>in</strong><br />

restoration (Table 6.1). In <strong>the</strong> fi eld of fl ood protection,<br />

recent major fl ood events <strong>in</strong> France, Germany and <strong>the</strong><br />

United K<strong>in</strong>gdom, for <strong>in</strong>stance, have accelerated <strong>the</strong> will<strong>in</strong>gness<br />

of authorities to enterta<strong>in</strong> catchment-oriented<br />

approaches and soft-eng<strong>in</strong>eer<strong>in</strong>g techniques of fl ood pro-<br />

tection, creat<strong>in</strong>g new opportunities for fl oodpla<strong>in</strong> restoration.<br />

The sheer cost of improv<strong>in</strong>g and ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g physical<br />

fl ood defences, <strong>in</strong> particular <strong>in</strong> rural areas, is rais<strong>in</strong>g <strong>in</strong>terest<br />

<strong>in</strong> alternative strategies. These alternative, <strong>in</strong>tegrated<br />

fl ood management strategies developed at <strong>the</strong> catchment<br />

scale are now considered viable to provide appropriate<br />

‘standards of service’ <strong>in</strong> areas of particular risk, while also<br />

ensur<strong>in</strong>g no net loss of ecosystem status, and even allow<strong>in</strong>g<br />

enhancement of aquatic, riparian and fl oodpla<strong>in</strong> environments.<br />

The evaluation of such strategies refl ects a<br />

policy shift from fl ood defence to fl ood risk management,<br />

and <strong>in</strong>cludes <strong>the</strong> possibility of <strong>in</strong>creas<strong>in</strong>g <strong>the</strong> frequency of<br />

fl ood<strong>in</strong>g and reduc<strong>in</strong>g <strong>the</strong> standard of service <strong>in</strong> some<br />

fl oodpla<strong>in</strong> locations where land is of relatively low value.<br />

This can have <strong>the</strong> effect of stor<strong>in</strong>g fl oodwater and attenuat<strong>in</strong>g<br />

hydrograph peaks, reduc<strong>in</strong>g fl ood potential <strong>in</strong> downstream<br />

high-value urban fl oodpla<strong>in</strong>s that are o<strong>the</strong>rwise at<br />

risk.<br />

Water protection agencies, concerned at water shortages<br />

and motivated by <strong>the</strong> EU Water Framework Directive,<br />

are also show<strong>in</strong>g <strong>in</strong>creased <strong>in</strong>terest <strong>in</strong> water fl ow regimes<br />

across whole catchments and <strong>in</strong> <strong>the</strong> potential of fl oodpla<strong>in</strong>s<br />

to improve water quality as part of a policy shift<br />

from downstream protection to upstream river bas<strong>in</strong> man-


Table 6.1 Recent policy shifts conducive to fl oodpla<strong>in</strong> restoration<br />

Policy fi eld Forces for change Policy issues<br />

agement. For nature conservationists restored fl oodpla<strong>in</strong>s<br />

represent important habitats that can contribute to meet<strong>in</strong>g<br />

biodiversity targets <strong>in</strong> accordance with <strong>the</strong> EU Habitats<br />

and Birds Directives; here <strong>the</strong> policy shift is <strong>in</strong> part from<br />

species protection to habitat enhancement. Political pressure<br />

is grow<strong>in</strong>g for more environmentally-sensitive forms<br />

of agriculture and forestry, creat<strong>in</strong>g new fund<strong>in</strong>g opportunities<br />

for extensive practices more suited to fl oodpla<strong>in</strong><br />

restoration, <strong>in</strong> a policy shift from agricultural support to<br />

<strong>in</strong>tegrated rural development. Agricultural policy focused<br />

on agri-environmental management means that fl oodpla<strong>in</strong>s<br />

orig<strong>in</strong>ally expensively dra<strong>in</strong>ed and protected can<br />

now be considered sites for re<strong>in</strong>stat<strong>in</strong>g o<strong>the</strong>r functions<br />

whose relative values are perceived to have <strong>in</strong>creased.<br />

F<strong>in</strong>ally, land-use plann<strong>in</strong>g regulations are be<strong>in</strong>g modifi ed<br />

to offer more effective protection of exist<strong>in</strong>g fl oodpla<strong>in</strong>s<br />

and, <strong>in</strong> some <strong>in</strong>stances, earmark<strong>in</strong>g land for <strong>the</strong> future<br />

restoration of fl oodpla<strong>in</strong>s. Spann<strong>in</strong>g <strong>the</strong>se sectoral policy<br />

shifts is a trend towards greater policy <strong>in</strong>tegration and<br />

stakeholder participation over schemes of this k<strong>in</strong>d (Pahl-<br />

Wostl, 2002, 2004), <strong>in</strong>formed <strong>in</strong> part by debates on susta<strong>in</strong>able<br />

development and new forms of governance<br />

(Bressers and Kuks, 2003).<br />

<strong>Manag<strong>in</strong>g</strong> <strong>the</strong>se changes is a challenge for <strong>the</strong> responsible<br />

agencies, which must grapple with questions of<br />

multi-functionality, multiple and nested scales, crosssectoral<br />

activity, policy <strong>in</strong>terplay, actor collaboration and<br />

issues of considerable complexity. One <strong>in</strong>stitutional framework<br />

for manag<strong>in</strong>g <strong>the</strong> changes <strong>in</strong> values, and <strong>the</strong> associated<br />

restoration of natural functions <strong>in</strong> fl oodpla<strong>in</strong>s, is<br />

provided by <strong>the</strong> Water Framework Directive <strong>in</strong> Europe,<br />

and by <strong>the</strong> tools developed for its implementation – <strong>River</strong><br />

Bas<strong>in</strong> Management Plans. In <strong>the</strong> United K<strong>in</strong>gdom, exist<strong>in</strong>g<br />

tools for strategic fl ood management (such as Catchment<br />

Flood Management Plans) now need adaptation to<br />

allow for ecosystem enhancement, comb<strong>in</strong><strong>in</strong>g <strong>the</strong> Envi-<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 81<br />

Flood protection Flood<strong>in</strong>g events; climate change;<br />

Risk management, soft eng<strong>in</strong>eer<strong>in</strong>g techniques,<br />

<strong>in</strong>frastructure costs; environmental quality natural fl ood storage<br />

Water protection EU Water Framework Directive; water<br />

Catchment-oriented approaches, fl ow regimes,<br />

quality/quantity problems<br />

wetlands, geomorphology<br />

Nature conservation EU Habitats Directive; concerns for<br />

biodiversity<br />

Functional fl oodpla<strong>in</strong> ecosystems<br />

Land-use plann<strong>in</strong>g L<strong>in</strong>kage of fl ood<strong>in</strong>g events to land use Plann<strong>in</strong>g mechanisms for protect<strong>in</strong>g and creat<strong>in</strong>g<br />

areas for fl ood retention<br />

Rural development EU Rural Development Regulation; spatial Integrated approaches to rural economic<br />

disparities<br />

development<br />

Agriculture Agenda 2000; public health concerns;<br />

environmental degradation<br />

Improved agri-environmental schemes<br />

ronment Agency’s roles <strong>in</strong> relation to fl ood management<br />

and conservation (and implementation of <strong>the</strong> EU Habitats<br />

Directive). These tools will thus necessarily focus on<br />

multi-functional management across a catchment-reachhabitat<br />

scale hierarchy, which will imply a variation <strong>in</strong> <strong>the</strong><br />

level of detailed plann<strong>in</strong>g at different scales. At <strong>the</strong> catchment<br />

scale, <strong>the</strong> plann<strong>in</strong>g is essentially a strategic assessment<br />

for manag<strong>in</strong>g priorities <strong>in</strong> relation to budgetary<br />

provision, while at <strong>the</strong> reach scale plann<strong>in</strong>g is focused on<br />

<strong>the</strong> implementation of specifi c policies. This has implications<br />

for <strong>the</strong> mean<strong>in</strong>g of uncerta<strong>in</strong>ty, s<strong>in</strong>ce qualitative<br />

general goals set at <strong>the</strong> catchment strategic scale may<br />

allow greater fl exibility, while more restrictive quantitative<br />

goals may be set at <strong>the</strong> reach scale. However, underly<strong>in</strong>g<br />

this is <strong>the</strong> governance requirement for stakeholder <strong>in</strong>volvement,<br />

expos<strong>in</strong>g <strong>the</strong> question of <strong>the</strong> accountability, <strong>in</strong> democratic<br />

societies, of those <strong>in</strong>stitutions responsible for<br />

goal-sett<strong>in</strong>g <strong>in</strong> relation to restoration <strong>in</strong>itiatives. This <strong>in</strong>troduces<br />

ano<strong>the</strong>r level of uncerta<strong>in</strong>ty.<br />

Decision makers must thus manage complexity and<br />

uncerta<strong>in</strong>ty, and cope <strong>in</strong> an adaptive manner with un<strong>in</strong>tended<br />

negative effects which emerge alongside <strong>the</strong> advantages<br />

derived from <strong>the</strong> policies <strong>the</strong>y enact and implement.<br />

In <strong>the</strong> context of fl oodpla<strong>in</strong> restoration, <strong>the</strong>y may seek to<br />

identify ‘reference’ conditions towards which <strong>the</strong> Water<br />

Framework Directive encourages a shift; however, <strong>the</strong>re is<br />

considerable uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> defi nition of such reference,<br />

or target, conditions. This ‘goal’ of a restoration<br />

<strong>in</strong>itiative may suggest identifi cation of a set of performance<br />

<strong>in</strong>dicators, but here aga<strong>in</strong> <strong>the</strong>re is uncerta<strong>in</strong>ty. If <strong>the</strong><br />

strategic goal of restoration is to recover <strong>the</strong> dynamics of<br />

natural process and function (by lett<strong>in</strong>g <strong>the</strong> river do <strong>the</strong><br />

work aga<strong>in</strong>), <strong>the</strong> performance criteria are very different,<br />

and necessarily less rigid given <strong>the</strong> time-variability of<br />

fl ows, than if <strong>the</strong> goal is set as a restoration of a specifi c<br />

(static) form.


82 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

There has thus been a grow<strong>in</strong>g <strong>in</strong>terest <strong>in</strong> restoration of<br />

riparian and fl oodpla<strong>in</strong> ecosystems, driven by <strong>in</strong>creas<strong>in</strong>g<br />

knowledge of <strong>the</strong> biophysical l<strong>in</strong>kages between parts of<br />

river bas<strong>in</strong>s and, <strong>in</strong> Europe, by shift<strong>in</strong>g policy directions<br />

and an <strong>in</strong>creased frequency and severity of fl ood events.<br />

Awareness of <strong>the</strong> potential impacts of global climate<br />

change and of <strong>the</strong> impacts of river management activities<br />

such as dam build<strong>in</strong>g on hydrological patterns <strong>in</strong> river<br />

bas<strong>in</strong>s (Montgomery and Boulton, 2003; World Commission<br />

on Dams, 2000) have led to a broaden<strong>in</strong>g of approaches<br />

to fl ood management. Coupled with a wider acceptance of<br />

multi-functional fl oodpla<strong>in</strong>s and more <strong>in</strong>tegrated policy<br />

contexts for <strong>the</strong>se, this has led to <strong>the</strong> proliferation of river<br />

restoration projects <strong>in</strong> many countries (Palmer et al., 2005).<br />

However, river restoration that <strong>in</strong>volves <strong>the</strong> planned return<br />

of fl ood<strong>in</strong>g and its associated geomorphological processes<br />

also reduces <strong>the</strong> predictability of river behaviour for river<br />

managers. The uncerta<strong>in</strong>ty associated with this development<br />

is <strong>the</strong> subject of this chapter, which cont<strong>in</strong>ues with a<br />

consideration of <strong>the</strong> spatial and temporal dynamics that<br />

underp<strong>in</strong> <strong>the</strong> natural function of rivers and <strong>the</strong>ir riparian<br />

and fl oodpla<strong>in</strong> zones.<br />

6.3 THE NATURAL FUNCTIONAL DYNAMICS<br />

OF RIPARIAN ENVIRONMENTS<br />

To achieve <strong>the</strong> goal of <strong>in</strong>tegrated management of fl oodpla<strong>in</strong>s,<br />

<strong>in</strong>clud<strong>in</strong>g restoration of <strong>the</strong>ir physical and ecological<br />

functions, it is necessary to understand some of <strong>the</strong><br />

key relationships that determ<strong>in</strong>e <strong>the</strong> health of <strong>the</strong> aquatic,<br />

riparian and fl oodpla<strong>in</strong> ecosystems. This section accord<strong>in</strong>gly<br />

briefl y reviews this understand<strong>in</strong>g, focus<strong>in</strong>g particularly<br />

on scale and spatial relationships with<strong>in</strong> <strong>the</strong> dra<strong>in</strong>age<br />

bas<strong>in</strong>, and on <strong>the</strong> dynamics of <strong>in</strong>ter-related processes that<br />

defi ne <strong>the</strong> functional status of rivers and <strong>the</strong>ir fl oodpla<strong>in</strong>s<br />

(Malanson, 1993).<br />

6.3.1 Scale and Spatial Relationships: Longitud<strong>in</strong>al<br />

and Lateral<br />

It is fi rst necessary to recognise that a river responds to<br />

conditions with<strong>in</strong> <strong>the</strong> upstream catchment dra<strong>in</strong><strong>in</strong>g to it,<br />

and that <strong>the</strong> river corridor is <strong>the</strong>refore <strong>the</strong> terrestrial low<br />

po<strong>in</strong>t which receives water fl ows, sediments, nutrients and<br />

plant propagules from its contribut<strong>in</strong>g area (Brierly and<br />

Fryirs, 2000). Accord<strong>in</strong>gly, <strong>the</strong> status of a river reach is<br />

dependent on its catchment environment. In a natural river<br />

reach, <strong>the</strong>re will be a diversity of habitats – pools, riffl es,<br />

gravel bars, po<strong>in</strong>t bars, steep banks, levees, side channels<br />

and chutes, abandoned channels, ox-bow lakes, back-<br />

swamps and fl oodpla<strong>in</strong>s. These habitats will be preferentially<br />

occupied by particular species of fauna and fl ora<br />

(see, for example, Marston et al., 1995; Richards et al.,<br />

2002). It is often <strong>the</strong> case that conservation focuses on<br />

particular species, but a key to general ecological health<br />

is <strong>the</strong> ma<strong>in</strong>tenance of habitats (Ward and Tockner, 2001).<br />

However, <strong>the</strong> habitats refl ect <strong>the</strong> behaviour of <strong>the</strong> river at<br />

<strong>the</strong> reach scale. For example, if a river meanders without<br />

constra<strong>in</strong>t, and <strong>in</strong>undates its fl oodpla<strong>in</strong> roughly once every<br />

1–5 years, it is likely that it will create and ma<strong>in</strong>ta<strong>in</strong> a<br />

natural diversity of habitat. Whe<strong>the</strong>r this behaviour occurs<br />

will depend on <strong>the</strong> way <strong>in</strong> which <strong>the</strong> catchment is occupied<br />

and managed, and delivers water and sediment to <strong>the</strong><br />

reach <strong>in</strong> question. Thus <strong>the</strong> crucial connections to understand<br />

are those between <strong>the</strong> hydrology, sediment supply<br />

and ecology of <strong>the</strong> contribut<strong>in</strong>g catchment area, and <strong>the</strong><br />

character of <strong>the</strong> river reach to which it dra<strong>in</strong>s. This results<br />

<strong>in</strong> a structure of longitud<strong>in</strong>al relationships and upstream–<br />

downstream connectivity, while lateral connections are<br />

also critical <strong>in</strong> l<strong>in</strong>k<strong>in</strong>g <strong>the</strong> aquatic (river) and terrestrial<br />

(fl oodpla<strong>in</strong>) environments across <strong>the</strong> riparian zone.<br />

The ecological status of a river reach is strongly dependent<br />

on longitud<strong>in</strong>al (downstream) connectivity, and this<br />

is refl ected <strong>in</strong> <strong>the</strong> river cont<strong>in</strong>uum concept (Vannote et al.,<br />

1980; Petts et al., 2000). This is based on <strong>the</strong> idea that<br />

rivers transport water and sediment downstream through<br />

a systematic cont<strong>in</strong>uum of conditions from steep, headwater<br />

reaches with coarse bed material, and shallow,<br />

tumbl<strong>in</strong>g fl ow, <strong>in</strong> narrow valleys, to more gently-slop<strong>in</strong>g<br />

lowland reaches with fi ne silty, sandy beds, deeper, slow<br />

fl ow and wide fl oodpla<strong>in</strong>s. Related to this, <strong>the</strong> ecology<br />

changes systematically from upstream to downstream<br />

reaches, and a cont<strong>in</strong>uum here refl ects <strong>the</strong> chang<strong>in</strong>g<br />

nature of biological processes. Upstream, woody debris is<br />

<strong>in</strong>troduced from <strong>the</strong> riparian vegetation and is ‘processed’<br />

by ‘shredders’ (<strong>in</strong>vertebrates that consume leaves and<br />

woody material) <strong>in</strong> <strong>the</strong> stream, while also supply<strong>in</strong>g nutrients<br />

and dissolved organic carbon. Fur<strong>the</strong>r downstream,<br />

<strong>the</strong>se products of upstream processes are used by o<strong>the</strong>r<br />

species, as when fi ne particulate organic matter is ei<strong>the</strong>r<br />

collected or fi ltered by species which are ‘ga<strong>the</strong>rers’. This<br />

implies that management which prevents upstream biological<br />

functions will have downstream effects because of<br />

<strong>the</strong> cont<strong>in</strong>uous relationships exist<strong>in</strong>g along <strong>the</strong> river<br />

course. Similarly, manag<strong>in</strong>g <strong>the</strong> river by <strong>in</strong>troduc<strong>in</strong>g dams<br />

and weirs will <strong>in</strong>terrupt <strong>the</strong> cont<strong>in</strong>uity of migration of<br />

species by <strong>in</strong>hibit<strong>in</strong>g upstream fi sh migration to spawn<strong>in</strong>g<br />

sites, and <strong>the</strong> downstream fl ow of both seeds and plant<br />

material from which vegetative reproduction may take<br />

place.<br />

There may be some debate about <strong>the</strong> details of <strong>the</strong><br />

‘river cont<strong>in</strong>uum’; an alternative is <strong>the</strong> process doma<strong>in</strong>s


concept (Montgomery, 1999), which argues that localscale<br />

geomorphological and biological processes determ<strong>in</strong>e<br />

<strong>the</strong> stream habitat, <strong>the</strong> disturbance regimes and <strong>the</strong><br />

species <strong>in</strong>teractions that <strong>in</strong>fl uence stream communities<br />

and biodiversity. However, for practical purposes, whe<strong>the</strong>r<br />

<strong>the</strong>re are longitud<strong>in</strong>al zones, or a longitud<strong>in</strong>al cont<strong>in</strong>uum,<br />

is less important than that <strong>the</strong>re are strong upstream–<br />

downstream relationships. Aquatic organisms have<br />

evolved to be adjusted to <strong>the</strong> most probable set of physical<br />

conditions aris<strong>in</strong>g from <strong>the</strong> fl uvial geomorphology and<br />

hydrology. Downstream reaches rely on carbon <strong>in</strong>puts<br />

from upstream, while middle reaches are sites of primary<br />

production (with clear water and less shad<strong>in</strong>g by<br />

vegetation).<br />

Longitud<strong>in</strong>al relationsips are supplemented by lateral<br />

connections (Ward and Stanford, 1995). The high biodiversity<br />

of river corridors is, <strong>in</strong> part, a refl ection of <strong>the</strong><br />

diversity of habitat at <strong>the</strong> marg<strong>in</strong> between two dist<strong>in</strong>ct<br />

ecosystems, those of <strong>the</strong> river and <strong>the</strong> fl oodpla<strong>in</strong>. The<br />

riparian zone is an aquatic–terrestrial ecotone, and is a<br />

zone of transition and exchange which benefi ts from, and<br />

regulates, <strong>the</strong> processes and functions of <strong>the</strong> ecosystems<br />

it connects. Interference with one <strong>in</strong>evitably affects <strong>the</strong><br />

o<strong>the</strong>r, and <strong>the</strong> ecotone is itself a high priority for conservation.<br />

When <strong>the</strong> channel is deepened and dredged, or when<br />

embankments are created, <strong>the</strong> lateral connectivity between<br />

river and fl oodpla<strong>in</strong> is disrupted. In some cases this<br />

restricts access to lateral channels, side-arms and ox-bow<br />

lakes, which convert to terrestrial status. The bi-directional<br />

connection between <strong>the</strong> river and groundwater is<br />

<strong>in</strong>hibited and fl oodpla<strong>in</strong> recharge is restricted. The potential<br />

for fi sh species to use fl oodpla<strong>in</strong> woodland as a fl oodperiod<br />

refuge is prevented and <strong>the</strong> fi sh population suffers.<br />

Bank-side vegetation provides shade, which regulates<br />

water temperature, and tree roots a diversity of microhabitats,<br />

which benefi t <strong>the</strong> aquatic fauna; it also supplies<br />

organic material to <strong>the</strong> stream. Removal of <strong>the</strong> vegetation<br />

removes <strong>the</strong>se functions and causes a loss of productivity,<br />

habitat and biodiversity.<br />

6.3.2 Dynamics: Variable Flows, Sediment Delivery<br />

and Channel Migration<br />

Aquatic and riparian species are adapted to and require<br />

variable fl ows (<strong>in</strong>clud<strong>in</strong>g overbank fl ows), and are sensitive<br />

to <strong>the</strong> tim<strong>in</strong>g of that variation. They also require different<br />

fl ows at various stages of <strong>the</strong>ir life cycles. For<br />

example, salmonidae require a stable gravel substrate<br />

when spawn<strong>in</strong>g, but at o<strong>the</strong>r times depend on fl ows<br />

which dilate <strong>the</strong> gravel and fl ush out fi ne sediment that<br />

has accumulated <strong>in</strong> <strong>the</strong> pore spaces. The preferred water<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 83<br />

depth, velocity and substrate for a given species will also<br />

change between <strong>the</strong> fry, juvenile and adult life stages, and<br />

this requires that <strong>the</strong>re is access to spatial variation of<br />

habitat, which provides suitable refuge locations dur<strong>in</strong>g<br />

high fl ows for <strong>in</strong>dividuals which cannot survive extreme<br />

conditions.<br />

Some riparian tree species, such as black poplar<br />

(Populus nigra) and willows, require occasional large<br />

fl oods to create new sedimentary surfaces for colonisation<br />

and regeneration, but also need a gradual recession<br />

dur<strong>in</strong>g <strong>the</strong> period of seedl<strong>in</strong>g establishment for <strong>the</strong>ir survival.<br />

The tim<strong>in</strong>g of high fl ow needs to match <strong>the</strong> release<br />

of seeds <strong>in</strong> late spr<strong>in</strong>g and early summer; this is an issue<br />

<strong>in</strong> river bas<strong>in</strong>s with reservoirs or water transfers that<br />

reduce fl ows and change <strong>the</strong> natural tim<strong>in</strong>g of high fl ows.<br />

The role of high fl ows is represented <strong>in</strong> <strong>the</strong> fl ood pulse<br />

concept (Junk et al., 1989; Middleton, 1999), which<br />

emphasises that fl oods cause channel migration and create<br />

new habitat, and supply nutrients and genetic material<br />

(seeds and plant material) to <strong>the</strong> riparian environment.<br />

However, <strong>the</strong>se processes can occur across a range of<br />

fl ows, so a general idea of a range of fl ow pulses (Tockner<br />

et al., 2000) is also found <strong>in</strong> <strong>the</strong> literature on riparian<br />

ecology.<br />

Of course, fl ows that are too high may both<br />

cause damage and create anaerobic conditions which<br />

prevent germ<strong>in</strong>ation and cause mortality <strong>in</strong> early seedl<strong>in</strong>gs.<br />

This illustrates that <strong>the</strong> process of manag<strong>in</strong>g<br />

fl ows for ecological benefi ts requires a subtle understand<strong>in</strong>g<br />

of <strong>the</strong> impacts of fl ows on <strong>the</strong> life histories of a<br />

range of species. This has given rise to <strong>the</strong> concept of<br />

‘environmental fl ows’, which <strong>in</strong>clude a wide range of fl ow<br />

levels at different times, and with different recurrences,<br />

each of which contributes to regeneration and to different<br />

life-cycle stages of different aspects of <strong>the</strong> aquatic and<br />

riparian ecology (Hill and Platts, 1991; Mahoney and<br />

Rood, 1998). The management of such variable fl ows,<br />

seasonally and <strong>in</strong>ter-annually, is a complex and uncerta<strong>in</strong><br />

process demand<strong>in</strong>g an adaptive response to <strong>the</strong> unfold<strong>in</strong>g<br />

of <strong>the</strong> unpredictable climatically-driven fl ow regime.<br />

However, it is a process which is now occurr<strong>in</strong>g <strong>in</strong> many<br />

climatic regions and is considered fur<strong>the</strong>r <strong>in</strong> <strong>the</strong> follow<strong>in</strong>g<br />

section.<br />

An important characteristic of natural, unmanaged<br />

rivers is that <strong>the</strong>y migrate across <strong>the</strong>ir fl oodpla<strong>in</strong>s at rates<br />

that depend on <strong>the</strong> energy of <strong>the</strong>ir fl ows and <strong>the</strong> resistance<br />

of <strong>the</strong>ir perimeter sediments to entra<strong>in</strong>ment, erosion and<br />

transport. This means that <strong>the</strong>y gradually turn over <strong>the</strong>ir<br />

fl oodpla<strong>in</strong> sediments, and that <strong>the</strong> fl oodpla<strong>in</strong> consists of a<br />

mosaic of surfaces of vary<strong>in</strong>g age and sedimentology<br />

giv<strong>in</strong>g rise to a mosaic of diverse habitats (Salo et al.,<br />

1986; Nilsson et al., 1991b).


84 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

The most well-known form of such migration is that of<br />

river meander<strong>in</strong>g, which arises because of bank erosion<br />

on <strong>the</strong> outside of bends and <strong>the</strong> deposition of po<strong>in</strong>t bars<br />

on <strong>the</strong> <strong>in</strong>sides. This process is what creates <strong>the</strong> diversity<br />

of habitat and is <strong>the</strong>refore very important for biodiversity.<br />

It is commonly understood that disturbance of ecosystems<br />

is critical for <strong>the</strong> structur<strong>in</strong>g and ma<strong>in</strong>tenance of biodiversity,<br />

as it stops a plant succession from progress<strong>in</strong>g to a<br />

uniform mature state and cont<strong>in</strong>ually resets <strong>the</strong> succession<br />

locally and re-<strong>in</strong>troduces pioneer species. In some ecosystems,<br />

disturbance may be caused by w<strong>in</strong>d, fi re or human<br />

<strong>in</strong>fl uences (<strong>in</strong> shift<strong>in</strong>g cultivation, for example). In river<br />

corridors, it is caused by erosion, deposition and river<br />

migration. The <strong>in</strong>termediate disturbance hypo<strong>the</strong>sis<br />

(Connell, 1978; Ward et al., 1999) argues that maximum<br />

biodiversity occurs under conditions of <strong>in</strong>termediate disturbance.<br />

Too much disturbance (as <strong>in</strong> very dynamic<br />

braided rivers) results <strong>in</strong> a preponderance of pioneer<br />

species, while too little disturbance allows a succession to<br />

progress towards a mature and relatively uniform plant<br />

community. Both have lower biodiversity than occurs<br />

with rates of disturbance that are <strong>in</strong>termediate. If <strong>in</strong> river<br />

corridors it is high fl ows, erosion, deposition and channel<br />

migration (<strong>the</strong> river dynamics) that promote biodiversity,<br />

it follows that practices of river management designed to<br />

<strong>in</strong>hibit movement of <strong>the</strong> river (such as embankment and<br />

bank protection) are likely to reduce biodiversity. Ecological<br />

status will improve <strong>the</strong>refore when rivers are given <strong>the</strong><br />

freedom to move. This is a concept which is now enshr<strong>in</strong>ed<br />

<strong>in</strong> policy for <strong>the</strong> management of <strong>the</strong> tributaries of <strong>the</strong><br />

Rh<strong>in</strong>e and <strong>the</strong> Meuse <strong>in</strong> The Ne<strong>the</strong>rlands, and <strong>in</strong>volves<br />

relocation of embankments, <strong>in</strong>creased off-channel fl ood<br />

storage and restoration of <strong>the</strong> functions of abandoned<br />

channels and side-arms.<br />

From this outl<strong>in</strong>e it is evident that <strong>the</strong> characteristics of<br />

<strong>the</strong> natural biophysical system of <strong>the</strong> river-fl oodpla<strong>in</strong><br />

environment are connectivity, spatial variability and temporal<br />

<strong>in</strong>stability (dynamics). It is generally acknowledged<br />

that restoration of <strong>the</strong>se characteristics is a necessary component<br />

of any restoration <strong>in</strong>itiative although <strong>the</strong> scale at<br />

which <strong>the</strong>y can be restored is very variable. The uncerta<strong>in</strong>ty<br />

posed by restor<strong>in</strong>g <strong>the</strong> very characteristics that were<br />

removed <strong>in</strong> order to make rivers more manageable and by<br />

extension less uncerta<strong>in</strong> <strong>in</strong> <strong>the</strong>ir behaviour also needs to<br />

be accommodated (see also Gregory and Downs, Chapter<br />

13). This is as much a question of chang<strong>in</strong>g attitudes<br />

(personal, public and <strong>in</strong>stitutional) through education as it<br />

is a question of understand<strong>in</strong>g more about <strong>the</strong> biophysical<br />

processes, although it rema<strong>in</strong>s <strong>the</strong> case that such understand<strong>in</strong>g<br />

is highly uncerta<strong>in</strong> when it requires prediction of<br />

future behaviour at a specifi c location with<strong>in</strong> a river<br />

system (e.g. Chapter 5).<br />

6.4 THE SCALE AND PRACTICE<br />

OF RESTORATION<br />

The practice of river restoration is <strong>in</strong>formed by <strong>the</strong> cont<strong>in</strong>u<strong>in</strong>g<br />

scientifi c <strong>in</strong>vestigation of <strong>the</strong> vital l<strong>in</strong>ks, at different<br />

spatial and temporal scales, between biotic and abiotic<br />

components across a catchment. However, it is also limited<br />

by compet<strong>in</strong>g needs for resources with<strong>in</strong> a catchment.<br />

The net result <strong>in</strong> many river bas<strong>in</strong>s is that river restoration<br />

takes place at relatively small, confi ned sites where re<strong>in</strong>statement<br />

of <strong>the</strong> l<strong>in</strong>ks between <strong>the</strong> river and its fl oodpla<strong>in</strong><br />

is limited <strong>in</strong> degree and extent and primarily focussed<br />

on re-establish<strong>in</strong>g lateral connectivity. It is important to<br />

dist<strong>in</strong>guish between river restoration that takes place at<br />

this spatially limited scale of <strong>the</strong> reach or section of a<br />

reach and river restoration that takes <strong>in</strong>to consideration <strong>the</strong><br />

longitud<strong>in</strong>al l<strong>in</strong>kages with<strong>in</strong> a river bas<strong>in</strong> and manages<br />

<strong>the</strong> primary <strong>in</strong>puts of water and sediment across <strong>the</strong> whole<br />

catchment. There are many thousands of river restoration<br />

projects worldwide that have been implemented at <strong>the</strong> fi rst<br />

scale but far fewer examples of resource management<br />

tak<strong>in</strong>g place at <strong>the</strong> scale of <strong>the</strong> catchment (see Chapter 13).<br />

At both scales <strong>the</strong>re is a considerable challenge for scientists<br />

to defi ne ecosystem needs <strong>in</strong> a way that can guide<br />

policy formulation and management action (Poff et al.,<br />

2003). However, <strong>the</strong> challenge is considerably greater at<br />

<strong>the</strong> scale of <strong>the</strong> catchment because levels of uncerta<strong>in</strong>ty<br />

about <strong>the</strong> ecological and o<strong>the</strong>r outcomes and <strong>the</strong> number<br />

and range of stakeholders that need to engage with <strong>the</strong><br />

process <strong>in</strong>crease rapidly as spatial scale <strong>in</strong>creases. Fur<strong>the</strong>rmore,<br />

until quite recently, <strong>the</strong> water needs of humans<br />

and those of ecosystems have been seen <strong>in</strong> competition<br />

(Richter et al., 2003) and <strong>the</strong> quantum leap <strong>in</strong> human<br />

perception from this to view<strong>in</strong>g ecosystems as legitimate<br />

users of water (K<strong>in</strong>g and Louwe, 1998; Naiman et al.,<br />

2002) has largely still to be made.<br />

To illustrate <strong>the</strong> issues and practice of work<strong>in</strong>g at<br />

different scales, <strong>the</strong> example of <strong>the</strong> needs of fl oodpla<strong>in</strong><br />

forest ecosystems and <strong>the</strong> different scales at which those<br />

needs can be provided is considered here. There are many<br />

<strong>in</strong>ter-related variables operat<strong>in</strong>g <strong>in</strong> a fully functional fl oodpla<strong>in</strong><br />

forest ecosystem but, never<strong>the</strong>less, <strong>the</strong>ir diverse,<br />

mobile vegetation mosaics can be said to have four essential<br />

requirements to be self-regenerat<strong>in</strong>g (Table 6.2). To<br />

get all of <strong>the</strong>se it is necessary to manage <strong>the</strong> disturbance<br />

processes that arrive at a fl oodpla<strong>in</strong> and this can be done<br />

<strong>in</strong> a number of ways at both <strong>the</strong> reach and catchment<br />

scales.<br />

6.4.1 Catchment-Scale Management<br />

At <strong>the</strong> scale of <strong>the</strong> catchment, restoration <strong>in</strong>itiatives are<br />

likely to <strong>in</strong>volve management of physical processes <strong>in</strong> one


or more places <strong>in</strong> <strong>the</strong> catchment upstream of <strong>the</strong> fl oodpla<strong>in</strong>,<br />

so that <strong>the</strong>y eventually have an effect on <strong>the</strong> disturbances<br />

arriv<strong>in</strong>g <strong>in</strong> <strong>the</strong> fl oodpla<strong>in</strong> zone. This type of<br />

disturbance management is ‘<strong>in</strong>direct’ but it is also <strong>the</strong><br />

most desirable for long term successful and self-susta<strong>in</strong><strong>in</strong>g<br />

restoration or management of fl oodpla<strong>in</strong> forests. It allows<br />

<strong>the</strong> river to fl ood and <strong>the</strong> channel to move and create its<br />

own sites for <strong>the</strong> regeneration of trees. It is, however, quite<br />

diffi cult to achieve for a variety of reasons:<br />

• It is not easy to predict where disturbances will have an<br />

<strong>in</strong>fl uence <strong>in</strong> <strong>the</strong> fl oodpla<strong>in</strong> zone and it is <strong>the</strong>refore an<br />

uncerta<strong>in</strong> form of management.<br />

• Unpredictability and reduced control have not been<br />

desirable features for river managers and as such <strong>the</strong>y<br />

tend to be avoided.<br />

• In highly managed and fragmented river systems <strong>the</strong>re<br />

may be too many human <strong>in</strong>terventions for it to work.<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 85<br />

Table 6.2 The four essential requirements for a self-regenerat<strong>in</strong>g fl oodpla<strong>in</strong> forest (from Hughes and Muller, 2003; Hughes et al.,<br />

2005)<br />

Requirement Rationale<br />

Flows needed by fl oodpla<strong>in</strong> forests • Regular fl ows which replenish and ma<strong>in</strong>ta<strong>in</strong> fl oodpla<strong>in</strong> water tables.<br />

These fl ows allow established trees to grow.<br />

• Periodic high fl ows which cause channel movement and sediment<br />

deposition. These provide potential regeneration sites and should be<br />

variable between years.<br />

• Well-timed fl ows through <strong>the</strong> fi rst grow<strong>in</strong>g season which allow<br />

delivery of seeds to <strong>the</strong> fl oodpla<strong>in</strong> and establishment of seedl<strong>in</strong>gs.<br />

Unseasonal high fl ows can cause high mortality to seedl<strong>in</strong>gs <strong>in</strong> <strong>the</strong>ir<br />

fi rst grow<strong>in</strong>g season.<br />

Regeneration sites needed by fl oodpla<strong>in</strong> forests • Open sites as many pioneer tree species typical of fl oodpla<strong>in</strong> forests<br />

cannot tolerate competition.<br />

• Sites that are moist through <strong>the</strong> fi rst grow<strong>in</strong>g season to facilitate<br />

regeneration.<br />

• Sites near <strong>the</strong> water’s edge because <strong>the</strong>se tend to be moister and catch<br />

organic debris. However, sites right on <strong>the</strong> water’s edge tend to suffer<br />

from fl ow disturbance and waterlogg<strong>in</strong>g.<br />

• A variety of sediment types to provide regeneration niches for a<br />

variety of species.<br />

Water table conditions needed by fl oodpla<strong>in</strong> forests • Water tables accessible to <strong>the</strong> roots of seedl<strong>in</strong>gs through <strong>the</strong>ir fi rst<br />

grow<strong>in</strong>g season.<br />

• Gradual recession of water tables follow<strong>in</strong>g a fl ood.<br />

• Limited waterlogg<strong>in</strong>g.<br />

Propagation materials needed by fl oodpla<strong>in</strong> forests • Seeds which are carried by <strong>the</strong> river and deposited dur<strong>in</strong>g fl oods. The<br />

phenology of seed release and <strong>the</strong> tim<strong>in</strong>g of fl ood peaks are critical <strong>in</strong><br />

any year for successful establishment of seedl<strong>in</strong>gs.<br />

• Vegetative material which arrives by fl ood or is deposited locally.<br />

• Seeds that are carried <strong>in</strong> <strong>the</strong> w<strong>in</strong>d. Whereas seeds carried <strong>in</strong> <strong>the</strong> river<br />

always move from upstream areas to downstream areas, seeds carried<br />

<strong>in</strong> <strong>the</strong> w<strong>in</strong>d tend to move <strong>in</strong> <strong>the</strong> direction of prevail<strong>in</strong>g w<strong>in</strong>ds.<br />

• It requires consensus among a huge number of<br />

stakeholders.<br />

The ways <strong>in</strong> which this catchment-scale management<br />

of water and sediment resources takes place is varied<br />

but <strong>the</strong>re are now a number of both established and<br />

emerg<strong>in</strong>g methodologies which have been put <strong>in</strong>to practice,<br />

many of which are reviewed more fully elsewhere<br />

(Arth<strong>in</strong>gton, 1998; Hughes and Rood, 2003; Postel and<br />

Richter, 2003):<br />

• Managed releases downstream from impoundments.<br />

This methodology is relatively simple <strong>in</strong> concept and<br />

<strong>in</strong>volves plann<strong>in</strong>g fl ow releases from structures such as<br />

dams so that <strong>the</strong>y provide maximum benefi t to downstream<br />

aquatic and riparian ecosystems as well as to<br />

o<strong>the</strong>r users. In practice this can only be contemplated<br />

where <strong>the</strong> eng<strong>in</strong>eer<strong>in</strong>g design of <strong>the</strong> dam structure gives


86 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

suffi cient control to release fl ows. The approach has<br />

been well-tried <strong>in</strong> North America and particularly good<br />

examples can be found <strong>in</strong> <strong>the</strong> St Mary <strong>River</strong> <strong>in</strong> Alberta,<br />

Canada (Rood and Mahoney, 2000) (Case Study 6.1)<br />

and <strong>in</strong> <strong>the</strong> Truckee <strong>River</strong> of Nevada, USA (Rood et al.,<br />

2003). Both examples use <strong>the</strong> Recruitment Box Model<br />

of Mahoney and Rood (1998) which allows quantitative<br />

description of <strong>the</strong> comb<strong>in</strong>ed requirement for appropriately<br />

timed high fl ows to create and saturate suitable<br />

fl oodpla<strong>in</strong> sites downstream and subsequent gradual<br />

fl ow recession (ramp<strong>in</strong>g rates) to permit seedl<strong>in</strong>g survival.<br />

Detailed knowledge of <strong>the</strong> requirements of <strong>the</strong><br />

germ<strong>in</strong>ation and seedl<strong>in</strong>g establishment phases of <strong>the</strong><br />

life-cycles of target tree species have to be known and<br />

ecologically relevant fl ows have to be characterised.<br />

There is considerable debate on how best to characterise<br />

<strong>the</strong> most ecologically critical aspects of fl ow regimes<br />

(Olden and Poff, 2003) such that an optimal balance is<br />

struck between a m<strong>in</strong>imum of hydrological <strong>in</strong>dices and<br />

maximum explanation of ecosystem function by <strong>the</strong>se<br />

<strong>in</strong>dices. The process becomes more diffi cult as projects<br />

go beyond prescrib<strong>in</strong>g fl ows for a s<strong>in</strong>gle ecosystem, like<br />

Case Study 6.1<br />

The ‘Recruitment Box’ model developed by Mahoney<br />

and Rood (1998) del<strong>in</strong>eates a zone on a fl oodpla<strong>in</strong>,<br />

defi ned by elevation and time, <strong>in</strong> which riparian cottonwood<br />

seedl<strong>in</strong>gs are likely to become successfully<br />

established if streamfl ow patterns are favourable<br />

(Figure 6.2). Along <strong>the</strong> St Mary <strong>River</strong> <strong>in</strong> Alberta,<br />

Canada, fl ow regulation from a headwater dam built <strong>in</strong><br />

1953 led to high mortality of established cottonwood<br />

(Populus deltoides) trees and no recruitment of new<br />

trees <strong>in</strong> <strong>the</strong> fl oodpla<strong>in</strong> downstream due to <strong>in</strong>suffi cient<br />

fl ows at critical times <strong>in</strong> <strong>the</strong> grow<strong>in</strong>g season (Rood<br />

et al., 1995) (Figure 6.3). Dur<strong>in</strong>g <strong>the</strong> 1990s, after identifi<br />

cation of <strong>the</strong> cause of <strong>the</strong> problem, regional water<br />

resource managers implemented changes <strong>in</strong> <strong>the</strong> operation<br />

of <strong>the</strong> St Mary Dam (Rood and Mahoney, 2000)<br />

(Figure 6.4). In particular, fl ows were designed to<br />

provide a gradual reduction <strong>in</strong> <strong>the</strong> fall<strong>in</strong>g limb of <strong>the</strong><br />

hydrograph after <strong>the</strong> spr<strong>in</strong>g snow-melt fl ood ra<strong>the</strong>r<br />

than <strong>the</strong> abrupt fall that occurred dur<strong>in</strong>g <strong>the</strong> post-dam<br />

period. This occurred because <strong>the</strong> dam operators shut<br />

<strong>the</strong> spillway gates at <strong>the</strong> dam to divert water <strong>in</strong>to irrigation<br />

channels. The gradual recession of fl oods was<br />

considered vital for replenishment of water tables <strong>in</strong><br />

<strong>the</strong> fl oodpla<strong>in</strong> zone and for <strong>the</strong> establishment of cottonwood<br />

seedl<strong>in</strong>gs whose roots are unable to ma<strong>in</strong>ta<strong>in</strong><br />

contact with <strong>the</strong> water table if it falls too rapidly.<br />

Figure 6.2 The recruitment Box model applied to <strong>the</strong> lower St<br />

Mary <strong>River</strong> (modifi ed from Mahoney and Rood, 1998)<br />

(a)<br />

(b)<br />

Figure 6.3 (a) A photograph of <strong>the</strong> St Mary <strong>River</strong> fl oodpla<strong>in</strong><br />

taken <strong>in</strong> 1991 shows dead cottonwoodtrees before successful<br />

fl ood releases were implemented (Photograph by Franc<strong>in</strong>e<br />

Hughes). (b) (See also colour plate section) By 2002 <strong>the</strong>re is<br />

signifi cant growth of young cottonwood trees on <strong>the</strong> fl oodpla<strong>in</strong><br />

follow<strong>in</strong>g planned releases (Photograph by Stewart Rood)


Figure 6.4 The two pairs of graphs depict a stage hydrograph<br />

with superimposed recruitment box at <strong>the</strong> ‘ideal’ time and elevation<br />

and with ideal drawdown rates. (a) In 1964, a post-dam fl ood<br />

was managed for maximal cut-back ra<strong>the</strong>r than naturalised recession.<br />

Regeneration of trees did not occur <strong>in</strong> that year. (b) In 1995,<br />

a managed fl ow, us<strong>in</strong>g <strong>the</strong> recruitment box as a guide, successfully<br />

promoted regeneration with a well-timed fl ood peak and<br />

suitable fl ood recession rates through <strong>the</strong> fi rst grow<strong>in</strong>g season<br />

(from Rood and Mahoney, 2000)<br />

a fl oodpla<strong>in</strong> forest, <strong>in</strong>to satisfy<strong>in</strong>g multiple ecosystem<br />

needs.<br />

• Flow allocation methodologies. <strong>River</strong> fl ows are altered<br />

as soon as water is used for human purposes. Richter<br />

et al. (2003) state that ‘<strong>the</strong> ultimate challenge of ecologically<br />

susta<strong>in</strong>able water management is to design and<br />

implement a water management program that stores and<br />

diverts water for human purposes <strong>in</strong> a manner that does<br />

Case Study 6.2<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 87<br />

not cause affected ecosystems to degrade or simplify’.<br />

There is necessarily a limit to how much water can be<br />

taken out of a river, and at what times of <strong>the</strong> year, if an<br />

ecosystem like a fl oodpla<strong>in</strong> forest is to rema<strong>in</strong> selfsusta<strong>in</strong>able.<br />

The question of how to allocate water to<br />

achieve susta<strong>in</strong>able water use for a range of ecosystems<br />

and human purposes is addressed by a series of holistic<br />

fl ow allocation methodologies, many of which have<br />

been developed <strong>in</strong> semi-arid areas <strong>in</strong> Australia, South<br />

Africa and North America. Exampes of <strong>the</strong>se models<br />

<strong>in</strong>clude <strong>the</strong> Bench Mark<strong>in</strong>g Methodology (Arth<strong>in</strong>gton,<br />

1998; Brizga, 2000) and <strong>the</strong> Flow <strong>Restoration</strong> Methodology<br />

(Arth<strong>in</strong>gton et al., 2000) <strong>in</strong> Australia; The Build<strong>in</strong>g<br />

Block Methodology (K<strong>in</strong>g and Louwe, 1998) and<br />

<strong>the</strong> DRIFT methodology (K<strong>in</strong>g et al., 2003; Brown and<br />

Joubert, 2003) <strong>in</strong> South Africa. Closely allied to <strong>the</strong>se<br />

models is <strong>the</strong> Adaptive Water Management Framework<br />

for <strong>in</strong>itiat<strong>in</strong>g ecologically susta<strong>in</strong>able water management<br />

programmes <strong>in</strong> <strong>the</strong> United States and elsewhere<br />

<strong>in</strong> <strong>the</strong> world by Richter et al. (2003). The general<br />

approach of <strong>the</strong>se methodologies is to determ<strong>in</strong>e <strong>the</strong><br />

‘environmental fl ows’ necessary to susta<strong>in</strong> aquatic and<br />

riparian ecosystems and <strong>the</strong>n to <strong>in</strong>tegrate <strong>the</strong> identifi ed<br />

fl ow needs with o<strong>the</strong>r fl ow requirements with<strong>in</strong> <strong>the</strong> river<br />

bas<strong>in</strong> on seasonal, <strong>in</strong>ter-annual and even decaded timescales.<br />

They are multi-stage, consensual or round-table<br />

approaches requir<strong>in</strong>g <strong>the</strong> <strong>in</strong>put of many experts and data<br />

on a range of aquatic and riparian ecosystems, on hydrological<br />

and geomorphological aspects, on modell<strong>in</strong>g of<br />

relationships between hydrological and biological attributes<br />

and eventually on all <strong>the</strong> o<strong>the</strong>r water uses <strong>in</strong> <strong>the</strong><br />

river bas<strong>in</strong>. It is at <strong>the</strong> stage when environmental fl ows<br />

are determ<strong>in</strong>ed that fl oodpla<strong>in</strong> forest requirements are<br />

<strong>in</strong>cluded <strong>in</strong> <strong>the</strong> process. The applicability of <strong>the</strong>se methodologies<br />

to <strong>the</strong> restoration of fl oodpla<strong>in</strong> forests <strong>in</strong><br />

Europe is discussed by Hughes and Rood (2003).<br />

Ano<strong>the</strong>r holistic approach to <strong>the</strong> management of water<br />

resources is <strong>the</strong> ‘alternative futures’ approach. This<br />

approach <strong>in</strong>cludes consideration of land uses across a<br />

catchment as well as fl ow allocation and among o<strong>the</strong>r<br />

places has been applied <strong>in</strong> <strong>the</strong> Willamette <strong>River</strong> Bas<strong>in</strong> <strong>in</strong><br />

Oregon, USA (Case Study 6.2).<br />

In <strong>the</strong> Willamette Bas<strong>in</strong>, <strong>in</strong> western Oregon, USA, an alternative futures analysis has been carried out to <strong>in</strong>form community<br />

decision mak<strong>in</strong>g regard<strong>in</strong>g land and water use <strong>in</strong> <strong>the</strong> river bas<strong>in</strong>. This is a participatory approach to river bas<strong>in</strong><br />

plann<strong>in</strong>g that presents stakeholders with a range of alternative scenarios <strong>in</strong>volv<strong>in</strong>g higher or lower levels of land and<br />

water use <strong>in</strong> <strong>the</strong> bas<strong>in</strong> and <strong>the</strong>ir resultant environmental impacts. In <strong>the</strong> Willamette Bas<strong>in</strong>, <strong>the</strong> current and historical<br />

landscapes were analysed and <strong>the</strong>n three future scenarios were generated, refl ect<strong>in</strong>g vary<strong>in</strong>g assumptions about land


88 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

and water use. Historical data on land and water use and population levels dat<strong>in</strong>g from 1850 were used and scenarios<br />

were projected to 2050 (Baker et al., 2004). Scenarios were evaluated on four areas of resource endpo<strong>in</strong>ts that were<br />

considered to be of value to stakeholders: water availability; river attributes such as channel structure, riparian and<br />

<strong>in</strong>stream ecosystem richness; ecological condition of streams (us<strong>in</strong>g fi sh and benthic <strong>in</strong>vertebrates as <strong>in</strong>dicators); and<br />

terrestrial wildlife <strong>in</strong> <strong>the</strong> bas<strong>in</strong> (Dole and Niemi, 2004) (Figure 6.5).<br />

The fl oodpla<strong>in</strong> of <strong>the</strong> Willamette <strong>River</strong> has been studied specifi cally with regard to prioritis<strong>in</strong>g parts of <strong>the</strong> historical<br />

fl oodpla<strong>in</strong> that might be particularly suitable for restoration us<strong>in</strong>g a landscape modell<strong>in</strong>g approach (Hulse<br />

and Gregory, 2004). This categorises <strong>the</strong> fl oodpla<strong>in</strong> <strong>in</strong>to areas of high and low potential for a range of both biophysical<br />

properties and socio-economic constra<strong>in</strong>ts considered at three spatial scales: <strong>the</strong> river network, <strong>the</strong> reach and<br />

<strong>the</strong> focal area. Scale was considered important because <strong>in</strong>teractions between bio-physical and socio-economic<br />

factors and with <strong>the</strong>m priorities, change at different spatial scales. Biophysical factors <strong>in</strong>cluded characteristics like<br />

channel complexity and hydrology and fl oodpla<strong>in</strong> vegetation type. Socio-economic constra<strong>in</strong>ts <strong>in</strong>cluded factors such<br />

as private or state ownership of fl oodpla<strong>in</strong> land and population density. The units used for apply<strong>in</strong>g <strong>the</strong>se categories<br />

are ‘slices’ of fl oodpla<strong>in</strong> at right angles to fl ow, each one kilometre long. Sections of <strong>the</strong> river with low constra<strong>in</strong>ts<br />

and high opportunity were identifi ed us<strong>in</strong>g this process and enabled prioritisation of candidate river and fl oodpla<strong>in</strong><br />

restoration sites. The results are presented as maps with marked areas that <strong>in</strong>tegrate representation of processes with<br />

patterns (Hulse and Gregory, 2004) (Figure 6.6). This methodology was used <strong>in</strong> <strong>the</strong> generation of scenarios for <strong>the</strong><br />

alternative futures analysis <strong>in</strong> <strong>the</strong> Willamette Bas<strong>in</strong> and particularly <strong>in</strong> assessment of <strong>the</strong> sensitivity of endpo<strong>in</strong>ts <strong>in</strong><br />

<strong>the</strong> river valley (Baker et al., 2004).<br />

Figure 6.5 Percentage change <strong>in</strong> <strong>the</strong> ‘forested riparian’ <strong>in</strong>dicator of natural resource condition <strong>in</strong> <strong>the</strong> Willamette <strong>River</strong> Bas<strong>in</strong>, <strong>in</strong> <strong>the</strong><br />

historical and three future scenarios. The <strong>in</strong>dicator is <strong>the</strong> percentage of a 120-metre wide riparian buffer strip with forest vegetation<br />

along all streams <strong>in</strong> <strong>the</strong> valley ecoregion (Hulse et al., 2002). The future scenario labels are described as: Conservative – a low level<br />

of development <strong>in</strong>volv<strong>in</strong>g a high degree of natural resource protection; Plan Trend – a level of development consistent with current<br />

policies and trends; Development – refl ects a loosen<strong>in</strong>g of current policies to allow a freer re<strong>in</strong> to market forces across all landscape<br />

components but still with<strong>in</strong> <strong>the</strong> range of what stakeholders considered plausible (From Ecological Application (2004) EA14-2,<br />

pp. 313–324, fi gure 4. Repr<strong>in</strong>ted with permission from The Ecological Society of America.)


<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 89<br />

Figure 6.6 Graphical example of river reaches (1, 2 and 3) with co<strong>in</strong>cident low constra<strong>in</strong>t and high opportunity to restore channel<br />

complexity and native fl oodpla<strong>in</strong> forest. The units of (a) Population density, are people per square kilometre circa 1990 with<strong>in</strong> each<br />

one kilometre slice of <strong>the</strong> fl oodpla<strong>in</strong>. The units of (b), number of structures, are rural build<strong>in</strong>gs per square kilometre circa 1990 with<strong>in</strong><br />

each one kilometre slice of <strong>the</strong> fl oodpla<strong>in</strong>. The units of (c), loss of channel complexity, are net <strong>in</strong>crease or decrease <strong>in</strong> channel length<br />

with<strong>in</strong> each one kilometre slice of <strong>the</strong> fl oodpla<strong>in</strong> between 1850 and 1995. The units of (d), loss of fl oodpla<strong>in</strong> forest, are net decrease<br />

<strong>in</strong> area of fl oodpla<strong>in</strong> forest with<strong>in</strong> each one kilometre slice of <strong>the</strong> fl oodpla<strong>in</strong> between 1850 and 1990. (From Hulse & Gregory (2004)<br />

Urban Ecosystems 7 (3): 295–314. Repr<strong>in</strong>ted with k<strong>in</strong>d permission from Spr<strong>in</strong>ger Science and Bus<strong>in</strong>ess Media.)


90 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Although <strong>the</strong> number of river bas<strong>in</strong>s to which <strong>the</strong>se<br />

emerg<strong>in</strong>g holistic methodologies have been applied is<br />

<strong>in</strong>creas<strong>in</strong>g, <strong>the</strong>y rema<strong>in</strong> <strong>the</strong> exception and not <strong>the</strong> rule.<br />

Richter and Postel (2004) give 350 as <strong>the</strong> number of river<br />

bas<strong>in</strong>s worldwide that are listed on a data base held by <strong>the</strong><br />

United States’ Nature Conservancy and are optimistic that<br />

obstacles to implement<strong>in</strong>g <strong>the</strong>se approaches are <strong>in</strong>creas<strong>in</strong>gly<br />

be<strong>in</strong>g overcome. Very positive steps have been taken<br />

<strong>in</strong> Europe to apply more holistic approaches to river bas<strong>in</strong><br />

management through <strong>the</strong> European Union Water Framework<br />

Directive and its required <strong>River</strong> Bas<strong>in</strong> Management<br />

Plans. However, management of total water volume and<br />

its seasonal distribution (currently carried out by most<br />

national river management agencies <strong>in</strong> Europe) tend to be<br />

limited to <strong>the</strong> management of water quantities <strong>in</strong> rivers to<br />

satisfy requirements for pollution dilution and <strong>the</strong> ma<strong>in</strong>tenance<br />

of specifi ed m<strong>in</strong>imum fl ows. This management is<br />

largely effected through control of water abstraction<br />

licences (Figure 6.7). It is usual for management of low<br />

fl ows to be separated from management of fl oods and this<br />

is an area that needs to be addressed to achieve a more<br />

holistic approach.<br />

• Sediment management. While a huge amount of literature<br />

has appeared on <strong>the</strong> management of fl ows, far less<br />

research has been carried out on <strong>the</strong> management of<br />

sediments and <strong>the</strong>re are also few examples of projects<br />

that <strong>in</strong>volve restoration of sediment loads <strong>in</strong> rivers. In<br />

<strong>the</strong> European context, sediment loads have drastically<br />

changed over <strong>the</strong> last 200 years <strong>in</strong> response to changes<br />

<strong>in</strong> mass movements <strong>in</strong> upper catchments, to sand and<br />

gravel extraction <strong>in</strong> fl oodpla<strong>in</strong> zones, to <strong>the</strong> <strong>in</strong>stallation<br />

of upstream impoundments and to <strong>the</strong> armour<strong>in</strong>g of<br />

river banks with artifi cial dykes. In <strong>the</strong> Drôme <strong>River</strong> <strong>in</strong><br />

France measures are proposed to restore sediment loads<br />

to <strong>the</strong> river. These <strong>in</strong>clude a moratorium on gravel<br />

extraction and clearance of river bank vegetation to remobilise<br />

sediment through bank erosion (Michelot,<br />

1995). The aim is to improve <strong>the</strong> delivery of sediment<br />

to <strong>the</strong> Ramieres du Val de Drôme Nature Reserve, which<br />

is designated for its high quality fl oodpla<strong>in</strong> forest <strong>in</strong> two<br />

active, braided river reaches.<br />

6.4.2 Reach-Scale Management<br />

At <strong>the</strong> scale of <strong>the</strong> reach, <strong>the</strong>re are two ma<strong>in</strong> approaches<br />

to carry<strong>in</strong>g out river and riparian restoration. The fi rst<br />

<strong>in</strong>volves manag<strong>in</strong>g physical processes locally <strong>in</strong> <strong>the</strong> fl oodpla<strong>in</strong><br />

zone, for example by <strong>in</strong>troduc<strong>in</strong>g sluices <strong>in</strong>to side<br />

channels to control water levels <strong>in</strong> selected parts of a<br />

fl oodpla<strong>in</strong>. The second <strong>in</strong>volves manag<strong>in</strong>g <strong>the</strong> landforms<br />

<strong>in</strong> <strong>the</strong> fl oodpla<strong>in</strong> so that <strong>the</strong> physical disturbances act differently<br />

on each part of <strong>the</strong> fl oodpla<strong>in</strong>. Both types of disturbance<br />

management are more directed to specifi ed<br />

reaches of a river than manag<strong>in</strong>g fl ows of water and sediment<br />

at a catchment scale. However, <strong>the</strong>y have <strong>the</strong> disadvantage<br />

that <strong>the</strong>y can only have a relatively local effect. In<br />

most parts of Europe it is <strong>the</strong> type of management that can<br />

most readily be promoted and <strong>in</strong> many river bas<strong>in</strong>s it has<br />

already taken place through a number of river restoration<br />

projects. <strong>Restoration</strong> at this scale is easier to achieve than<br />

catchment-scale management for a variety of reasons:<br />

Figure 6.7 A typical environmental allocation for a UK catchment is shown <strong>in</strong> this graph (from Environment Agency, 2002). It<br />

varies considerably through <strong>the</strong> seasons and is determ<strong>in</strong>ed through a technical approach called <strong>the</strong> Resource Assessment and Management<br />

(RAM) framework. First each river reach <strong>in</strong> <strong>the</strong> catchment is assessed on <strong>the</strong> basis of its physical characteristics, fi sh populations,<br />

macrophytes and macro-<strong>in</strong>vertebrates to produce an environmental weight<strong>in</strong>g. Five environmental weight<strong>in</strong>g bands are used to classsify<br />

<strong>the</strong> sensitivity of each reach to <strong>the</strong> effects of water abstraction. The environmental weight<strong>in</strong>g is used with long term fl ow duration<br />

data to derive an Ecological Flow Objective and <strong>the</strong> percentage left for abstraction. The Ecological Flow Objective seeks to protect<br />

low fl ows and fl ow variability by allow<strong>in</strong>g percentages of fl ow bands to be abstracted. The impacts of groundwater abstraction (both<br />

seasonally and spatially) on river fl ows is also <strong>in</strong>corporated <strong>in</strong>to <strong>the</strong> process. To preserve water levels <strong>in</strong> fl oodpla<strong>in</strong> sites adjacent to<br />

rivers, a site-scale approach has been used <strong>in</strong> parts of <strong>the</strong> UK us<strong>in</strong>g Water Level Management Plans. These are usually applied to sites<br />

designated as wetland nature reserves whose water tables are related to those of adjacent river courses <strong>in</strong> ei<strong>the</strong>r a direct or <strong>in</strong>direct<br />

way


• It is more predictable than catchment-scale management<br />

of disturbances and as such is relatively easy to control<br />

and poses less uncerta<strong>in</strong>ties.<br />

• <strong>River</strong> managers are already used to manag<strong>in</strong>g rivers at<br />

a local scale through <strong>the</strong> eng<strong>in</strong>eer<strong>in</strong>g of fl ood defences.<br />

• It can neatly be fi tted <strong>in</strong>to chosen river sections between<br />

reaches where o<strong>the</strong>r human <strong>in</strong>terventions dom<strong>in</strong>ate.<br />

• It requires consensus among a much smaller group of<br />

stakeholders.<br />

Exampes of restoration at this scale are now common<br />

and can be categorised <strong>in</strong>to a series of activities which may<br />

be carried out s<strong>in</strong>gly or <strong>in</strong> comb<strong>in</strong>ation at any <strong>in</strong>dividual<br />

site (Hughes and Muller, 2003):<br />

• <strong>Restoration</strong> of <strong>the</strong> channel wetted perimeter. These projects<br />

have as a ma<strong>in</strong> aim <strong>the</strong> improvement of <strong>in</strong>stream<br />

habitats, often for fi sh populations. They emphasise<br />

<strong>in</strong>creas<strong>in</strong>g <strong>the</strong> heterogeneity of physical habitat. For<br />

example, <strong>the</strong> Ecological recovery of <strong>the</strong> V<strong>in</strong>del and Pitte<br />

<strong>River</strong>s (EVP) project, <strong>in</strong> nor<strong>the</strong>rn Sweden <strong>in</strong>volves<br />

removal of <strong>in</strong>stream structures <strong>in</strong>stalled dur<strong>in</strong>g <strong>the</strong> n<strong>in</strong>eteenth<br />

century for driv<strong>in</strong>g logs for <strong>the</strong> timber <strong>in</strong>dustry.<br />

In addition, large boulders that were removed for smooth<br />

fl oat<strong>in</strong>g of log rafts are be<strong>in</strong>g put back to diversify<br />

<strong>in</strong>stream habitats (Nilsson et al., 2005).<br />

• Re-connection of side arms <strong>in</strong> rivers. Most projects <strong>in</strong><br />

this category aim to remove one or more sections of<br />

artifi cial embankments to allow fl ows to penetrate fl oodpla<strong>in</strong><br />

zones that have been cut off. There are a number<br />

of examples of this approach that are well documented<br />

<strong>in</strong> <strong>the</strong> literature, such as <strong>the</strong> Regelsbrunner Au project<br />

on <strong>the</strong> <strong>River</strong> Danube <strong>in</strong> Austria (Scheimer et al., 1999)<br />

or <strong>the</strong> L’Ile de la Platière and Rosillon Channel sites on<br />

<strong>the</strong> <strong>River</strong> Rhône <strong>in</strong> France (Michelot, 1995; Downs<br />

et al., 2002). These projects had as a major objective <strong>the</strong><br />

restoration of both <strong>the</strong> quality of fl oodpla<strong>in</strong> forests and<br />

improved opportunites for forest regeneration by<br />

<strong>in</strong>creas<strong>in</strong>g lateral connectivities between <strong>the</strong> channel<br />

and <strong>the</strong> fl oodpla<strong>in</strong>.<br />

• Increase <strong>in</strong> fl oodpla<strong>in</strong> storage capacity through sett<strong>in</strong>g<br />

back defences, lower<strong>in</strong>g fl ood defences or lower<strong>in</strong>g <strong>the</strong><br />

fl oodpla<strong>in</strong>. There are major and well-documented plans<br />

for a range of <strong>the</strong>se activities <strong>in</strong> <strong>the</strong> distributaries of <strong>the</strong><br />

<strong>River</strong> Rh<strong>in</strong>e <strong>in</strong> The Ne<strong>the</strong>rlands and also <strong>in</strong> <strong>the</strong> <strong>River</strong><br />

Meuse. They form part of a master plan to <strong>in</strong>crease fl ood<br />

storage capacity <strong>in</strong> <strong>the</strong>se rivers <strong>in</strong> anticipation of sealevel<br />

rise and <strong>in</strong>creas<strong>in</strong>g frequency of fl oods from<br />

upstream (Middelkoop and van Haselen, 1999). These<br />

activities are currently restricted to a series of discrete<br />

sites, such as at <strong>the</strong> Mill<strong>in</strong>gerwaard Nature Reserve <strong>in</strong><br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 91<br />

The Ne<strong>the</strong>rlands where fl ood defences have been<br />

lowered. The general approach <strong>in</strong> The Ne<strong>the</strong>rlands is<br />

shown <strong>in</strong> Figure 6.8. In <strong>the</strong> United K<strong>in</strong>gdom fl ood<br />

storage washlands are proposed along some rivers but<br />

<strong>in</strong> all <strong>the</strong>se projects <strong>the</strong> design of dra<strong>in</strong>age dykes and<br />

control structures will be very important <strong>in</strong> determ<strong>in</strong>g<br />

<strong>the</strong> amount of control on <strong>the</strong> length of time and depth<br />

that water stays <strong>in</strong> <strong>the</strong> washland. Creative confi guration<br />

of <strong>the</strong> fl oodpla<strong>in</strong> surface can mimic natural fl oodpla<strong>in</strong><br />

habitat heterogeneity to give conservation ga<strong>in</strong>s as well<br />

as fl ood storage ga<strong>in</strong>s if water control is fl exible enough<br />

to operate for <strong>the</strong> good of wetland ecosystems as well<br />

as for fl ood control.<br />

• Management of <strong>the</strong> river’s sediment load. Sedimentation<br />

is an essential process along <strong>the</strong> marg<strong>in</strong>s of river channels<br />

as newly created alluvial bars are prime regeneration<br />

sites for many species of fl oodpla<strong>in</strong> vegetation.<br />

Groynes can be used to create artifi cial ‘beaches’<br />

although <strong>the</strong>ir primary purpose is usually to ma<strong>in</strong>ta<strong>in</strong> a<br />

channel for navigation. Re-activation of erosion <strong>in</strong> sites<br />

where embankments have been removed will alter <strong>the</strong><br />

sediment loads downstream.<br />

• Management of riparian vegetation. This can take <strong>the</strong><br />

form of plant<strong>in</strong>g fl oodpla<strong>in</strong> forests or wait<strong>in</strong>g for natural<br />

regeneration to take place. In ei<strong>the</strong>r case, management<br />

of <strong>the</strong> vegetation that grows may be considered necessary,<br />

though this is not <strong>in</strong> <strong>the</strong> spirit of self-susta<strong>in</strong>ability.<br />

In many river bas<strong>in</strong>s, graz<strong>in</strong>g by domestic animals <strong>in</strong><br />

riparian zones prevents natural regeneration and fenc<strong>in</strong>g<br />

Figure 6.8 In The Ne<strong>the</strong>rlands, large parts of <strong>the</strong> Rh<strong>in</strong>e fl oodpla<strong>in</strong><br />

will be lowered as part of <strong>the</strong> Flood Action Plans drawn up<br />

by <strong>the</strong> International Rh<strong>in</strong>e Committee. The aim is to <strong>in</strong>crease<br />

fl ood storage capacity of <strong>the</strong> Rh<strong>in</strong>e fl oodpla<strong>in</strong>s. The proposed<br />

works are shown schematically <strong>in</strong> this diagram (from Middelkoop<br />

and Van Haselen, 1999)


92 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

off <strong>the</strong> riparian zone becomes an important form of<br />

management. Management of <strong>the</strong> vegetation also has<br />

implications for <strong>the</strong> volume of woody debris that arrives<br />

<strong>in</strong> a river. MacNally et al. (2002) suggest that <strong>in</strong> <strong>the</strong><br />

Murray-Darl<strong>in</strong>g <strong>River</strong> system <strong>in</strong> Australia only 15% of<br />

<strong>the</strong> former woody debris load is now present. There are<br />

projects where woody debris has been put back <strong>in</strong>to<br />

rivers and some where artifi cial log jams have been built<br />

to test <strong>the</strong>ir restoration effect (Dewberry et al., 1998).<br />

6.4.3 The Use of Reference Systems <strong>in</strong> <strong>River</strong><br />

<strong>Restoration</strong> Projects<br />

It is common to set objectives for river restoration projects<br />

by us<strong>in</strong>g reference systems. In <strong>the</strong> EU Water Framework<br />

Directive (2000/60/EC) it is a major and perhaps unrealistic<br />

aim to use a network of near-natural reference systems<br />

so that <strong>the</strong> ecological status of rivers can be measured<br />

aga<strong>in</strong>st <strong>the</strong>m. In essence this <strong>in</strong>volves fi nd<strong>in</strong>g a river<br />

system that has <strong>the</strong> attributes considered desirable <strong>in</strong> <strong>the</strong><br />

restored system and us<strong>in</strong>g it as a template on which to base<br />

<strong>the</strong> restoration activities. Criteria specifi cally for establish<strong>in</strong>g<br />

riparian reference conditions are proposed by Harris<br />

(1999) us<strong>in</strong>g multivariate analyses of aspects of vegetation<br />

community composition and structure. Typically a reference<br />

system will be part of or <strong>the</strong> whole of a less-damaged<br />

river system, preferably located <strong>in</strong> a similar bioclimatic<br />

zone and <strong>in</strong> a river bas<strong>in</strong> exhibit<strong>in</strong>g similar physiographic<br />

characteristics. Alternatively it can be an historic system,<br />

whose attributes are known from maps, old photographs<br />

or written accounts. A critique of <strong>the</strong> use of reference<br />

systems to defi ne objectives <strong>in</strong> river restoration is given <strong>in</strong><br />

Hughes et al. (2005), where <strong>the</strong> follow<strong>in</strong>g categories of<br />

problems encountered with <strong>the</strong> use of reference systems<br />

are discussed:<br />

• There are often no appropriate reference systems to<br />

use.<br />

• Many catchment parameters have changed s<strong>in</strong>ce <strong>the</strong><br />

times of chosen historic reference systems.<br />

• Climate change has been cont<strong>in</strong>uous through <strong>the</strong><br />

Holocene.<br />

• Projected climate change is of uncerta<strong>in</strong> magnitude.<br />

• Alien species have become common <strong>in</strong> <strong>the</strong> landscape<br />

and cannot be avoided.<br />

• Landscape context changes through time.<br />

Us<strong>in</strong>g reference systems can give river managers a misplaced<br />

confi dence <strong>in</strong> <strong>the</strong> predictability of ecological outcomes<br />

<strong>in</strong> river restoration projects (Hughes et al., 2005),<br />

although <strong>the</strong> degree to which <strong>the</strong> project outcomes and <strong>the</strong><br />

reference system co<strong>in</strong>cide can largely be decided at <strong>the</strong><br />

outset when objectives are set (Simons and Boeters, 1998).<br />

Evaluation of restoration projects is also usually carried<br />

out aga<strong>in</strong>st <strong>the</strong> reference system, although with highly<br />

variable levels of rigour (Anderson and Dugger, 1998;<br />

Stream Corridor Work<strong>in</strong>g Group, 1998).<br />

Never<strong>the</strong>less, reference systems can provide useful,<br />

broad guid<strong>in</strong>g images for restoration. In <strong>the</strong> United<br />

K<strong>in</strong>gdom, a number of sources of <strong>in</strong>formation are used.<br />

At <strong>the</strong> catchment scale, <strong>the</strong> National Vegetation Classifi cation<br />

(Rodwell, 1991a, 1991b, 1992, 1995, 2000), The<br />

National Biodiversity Network (NBN), <strong>the</strong> Multi-Agency<br />

Geographic Information for <strong>the</strong> Countryside (MAGIC)<br />

databases on Habitats and Sites of Special Scientifi c <strong>in</strong>terest<br />

(SSSI’s), provide <strong>in</strong>formation on <strong>the</strong> distribution of<br />

habitats, communities and species across <strong>the</strong> United<br />

K<strong>in</strong>gdom. For river reaches, <strong>the</strong>re may be <strong>River</strong> Morphology,<br />

<strong>River</strong> Habitat and <strong>River</strong> Corridor Surveys, and<br />

Hydro-Morphology Quality Assessments that have already<br />

been undertaken <strong>in</strong> some relatively undisturbed reaches<br />

(RSPB, NRA, RSNC, 1994). Us<strong>in</strong>g data on <strong>the</strong> catchment<br />

area, slope and river corridor width of <strong>the</strong> reaches, it may<br />

be possible to extrapolate to o<strong>the</strong>r river reaches with<br />

similar characteristics. Overall, <strong>the</strong> aim should be to<br />

develop a sense of how <strong>the</strong> spatial structure of <strong>the</strong> catchment<br />

hydrology and river morphology and ecology are<br />

related, and how <strong>the</strong> catchment and river function under<br />

‘natural’ conditions, bear<strong>in</strong>g <strong>in</strong> m<strong>in</strong>d that any system is <strong>in</strong><br />

a transient state over time.<br />

6.5 MONITORING AND EVALUATING<br />

RESTORATION<br />

6.5.1 What is Ecological Success and How Do We<br />

Evaluate It?<br />

A widely applicable scheme for evaluat<strong>in</strong>g <strong>the</strong> ecological<br />

success of river restoration projects does not currently<br />

exist, though many authors have emphasised <strong>the</strong> need for<br />

such a scheme (NRC, 1992; Downs et al., 2002). Part of<br />

<strong>the</strong> problem is <strong>in</strong> defi n<strong>in</strong>g exactly what is meant by ecological<br />

success and recently an attempt has been made to<br />

identify fi ve criteria that can be used for its measurement<br />

(Palmer et al., 2005):<br />

1. The design of an ecological river restoration project<br />

should be based on a specifi ed guid<strong>in</strong>g image of a more<br />

dynamic, healthy river that could exist at <strong>the</strong> site.<br />

2. The river’s ecological condition must be measurably<br />

improved.<br />

3. The river system must be more self-susta<strong>in</strong><strong>in</strong>g and<br />

resilient to external perturbations so that only m<strong>in</strong>imal<br />

follow-up ma<strong>in</strong>tenance is needed.


4. Dur<strong>in</strong>g <strong>the</strong> construction phase, no last<strong>in</strong>g harm is<br />

<strong>in</strong>fl icted on <strong>the</strong> ecosystem.<br />

5. Both pre- and post-construction ecological assessment<br />

is carried out and <strong>the</strong> <strong>in</strong>formation made available.<br />

Refi nements to this scheme have been proposed by<br />

Jansson et al. (2005) and Gillilan et al. (2005), <strong>in</strong>clud<strong>in</strong>g<br />

<strong>the</strong> need to <strong>in</strong>clude stated ecological mechanisms by<br />

which any <strong>in</strong>tended restoration will reach its goal. A major<br />

diffi culty with this scheme is giv<strong>in</strong>g practical mean<strong>in</strong>g to<br />

<strong>the</strong> concepts of resiliency and self-susta<strong>in</strong>ability, pr<strong>in</strong>cipally<br />

because <strong>the</strong>se are entirely relative terms that are also<br />

closely l<strong>in</strong>ked to <strong>the</strong> concept of dynamic equilibrium.<br />

The term ‘resilience’ was fi rst co<strong>in</strong>ed <strong>in</strong> <strong>the</strong> 1970s <strong>in</strong><br />

<strong>the</strong> fi eld of population ecology and used to characterise<br />

<strong>the</strong> magnitude of population perturbations a system could<br />

tolerate before chang<strong>in</strong>g <strong>in</strong>to some qualitatively different<br />

dynamic state (Holl<strong>in</strong>g, 1973; May, 1976a). More recently<br />

it has been used to describe <strong>the</strong> ability of an ecosystem to<br />

rega<strong>in</strong> a functional state follow<strong>in</strong>g disturbance and <strong>the</strong><br />

rapidity of this process is a measure of its resilience<br />

(War<strong>in</strong>g, 1989). The disturbance is usually temporary and<br />

if it is suffi ciently regular, component species of <strong>the</strong> ecosystem<br />

often evolve a dependence on it to complete <strong>the</strong>ir<br />

life cycles. Both ecological and evolutionary time scales<br />

must be considered <strong>in</strong> assess<strong>in</strong>g <strong>the</strong> signifi cance of disturbances<br />

of different magnitudes and frequencies <strong>in</strong> river<br />

and riparian ecosystems (Poff, 1992; Hughes, 1994; Dodds<br />

et al., 2004). As described earlier <strong>in</strong> this chapter, <strong>in</strong> <strong>the</strong><br />

case of riparian and fl oodpla<strong>in</strong> ecosystems, <strong>the</strong> disturbance<br />

is provided by fl oods and many riparian species<br />

have <strong>in</strong>deed evolved a dependence on <strong>the</strong>se fl oods for <strong>the</strong><br />

regeneration phase of <strong>the</strong>ir life cycles. ‘Dynamic equilibrium’<br />

encompasses <strong>the</strong> notion of chang<strong>in</strong>g parameters<br />

with<strong>in</strong> a stable framework and has often been used <strong>in</strong><br />

descriptions of whole ecosystems. If <strong>the</strong> stable framework<br />

changes <strong>the</strong>n a qualitatively different dynamic state aga<strong>in</strong><br />

prevails. The defi nition of a stable framework is usually<br />

determ<strong>in</strong>ed by <strong>the</strong> temporal and spatial scales of an <strong>in</strong>vestigation<br />

and often limited by an <strong>in</strong>vestigator’s consideration<br />

of change through time and over space. However, as<br />

stated by May (1976b), <strong>in</strong> <strong>the</strong> real world <strong>the</strong>re are no fi xed<br />

parameter values; environmental parameters and <strong>in</strong>teractions<br />

between organisms and between organisms and <strong>the</strong>ir<br />

environment are constantly fl uctuat<strong>in</strong>g. It follows that<br />

stable frameworks do not exist <strong>in</strong> <strong>the</strong> real world unless<br />

<strong>the</strong>y are artifi cially delimited spatially, temporally or<br />

both.<br />

Measur<strong>in</strong>g resilience to evaluate how successful we<br />

have been <strong>in</strong> return<strong>in</strong>g it to a riparian or fl oodpla<strong>in</strong> ecosystem<br />

dur<strong>in</strong>g a river restoration project is very diffi cult<br />

to do. If a framework has been identifi ed to describe a<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 93<br />

dynamic equilibrium, <strong>the</strong>n with<strong>in</strong> <strong>the</strong> defi ned framework<br />

it is possible to predict and measure recovery of a disturbed<br />

ecosystem. However, it is a huge challenge to<br />

predict <strong>the</strong> functions of ecosystems when <strong>the</strong>y beg<strong>in</strong> to<br />

fl uctuate along new trajectories caused by environmental<br />

change or <strong>in</strong>clude new species arrivals. S<strong>in</strong>ce all ecosystems,<br />

<strong>in</strong>clud<strong>in</strong>g restored ecosystems, are mov<strong>in</strong>g along<br />

fl uctuat<strong>in</strong>g trajectories, it follows that, conceptually, it is<br />

impossible to measure resilience and <strong>the</strong>refore to evaluate<br />

ecological success us<strong>in</strong>g this criterion (see Chapters 8 and<br />

11 for fur<strong>the</strong>r discussion). In practical terms it is possible<br />

to del<strong>in</strong>eate a framework for measur<strong>in</strong>g resilience though<br />

different types of ecologists might defi ne very different<br />

frameworks for <strong>the</strong> same restoration project. A fl oodpla<strong>in</strong><br />

forest ecologist might defi ne a framework where selfsusta<strong>in</strong>ability<br />

is measured <strong>in</strong> terms of turnover rates<br />

for fl oodpla<strong>in</strong> habitats (10 2 –10 3 years) related to fl ood<br />

return periods (Mahoney and Rood, 1998; Hughes and<br />

Rood, 2001). A fi sheries ecologist might assess selfsusta<strong>in</strong>ability<br />

<strong>in</strong> terms of provision of <strong>in</strong>stream habitats<br />

that permit completion of fi sh life cycles over annual or<br />

10 1 year time frames (e.g. Frissel and Nawa, 1992).<br />

The evaluation of success us<strong>in</strong>g such measurements<br />

also changes with scale and is discussed with respect to<br />

fl oodpla<strong>in</strong> forests by Hughes et al. (2005). When viewed<br />

at a small spatial scale (10 2 metres), habitat patches <strong>in</strong> <strong>the</strong><br />

forest will change from year to year as a response to<br />

channel movement and species arrival, death or migration.<br />

At a broader spatial scale of a reach or whole fl oodpla<strong>in</strong>,<br />

variability becomes less pronounced because <strong>the</strong> balance<br />

of different habitat patches rema<strong>in</strong>s more constant (Figures<br />

6.9(a) and 6.9(b)), particularly over short time frames of<br />

10 1 to 10 2 years (see Ward et al., 2002). However, over<br />

longer periods (10 3 years), this balance at <strong>the</strong> reach scale<br />

might shift <strong>in</strong> response to climate change, sea level change,<br />

isostatic uplift, change <strong>in</strong> availability of propagules or<br />

changes <strong>in</strong> <strong>the</strong> biophysical attributes of <strong>the</strong> catchment<br />

(Figure 6.9(c)). In this scenario, shift<strong>in</strong>g geomorphological<br />

processes (for example from aggradation to downcutt<strong>in</strong>g<br />

of <strong>the</strong> river valley) and changed channel patterns<br />

cause major changes to ecological vectors and patterns <strong>in</strong><br />

<strong>the</strong> fl oodpla<strong>in</strong>. Evaluation of success can only be relative<br />

to <strong>the</strong>se chang<strong>in</strong>g frameworks or else with<strong>in</strong> <strong>the</strong> context<br />

of a framework that has been held stable.<br />

6.5.2 What is Ecological Quality and How Does it<br />

Relate to Ecological Success?<br />

Closely related to <strong>the</strong> consideration of how to measure<br />

ecological success is <strong>the</strong> need to give mean<strong>in</strong>g to <strong>the</strong><br />

concept of ecological quality. Both concepts are dependent<br />

on <strong>the</strong> monitor<strong>in</strong>g or surveillance of a series of physical


94 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Figure 6.9 At <strong>the</strong> scale of a whole fl oodpla<strong>in</strong>, progressive or rapid changes can take place <strong>in</strong> <strong>the</strong> distribution of fl oodpla<strong>in</strong> vegetation<br />

communities. Thus <strong>in</strong> 6.9(a), t1 is biodiversity at an <strong>in</strong>itial time period and consists of vegetation communities a and b, present<br />

<strong>in</strong> <strong>the</strong> proportion of 4a to 3b. In 6.9(b), t2 is biodiversity at a later time. Vegetation communities a and b are still present <strong>in</strong> <strong>the</strong> same<br />

balance but <strong>the</strong>y are all <strong>in</strong> different locations follow<strong>in</strong>g shifts <strong>in</strong> channel location. Over much longer time frames (or over rapid time<br />

frames follow<strong>in</strong>g extreme events or human <strong>in</strong>tervention), <strong>the</strong>re may be changes <strong>in</strong> catchment parameters that alter <strong>the</strong> geomorphological<br />

patterns and hydrological activity of <strong>the</strong> river. In 6.9(c), <strong>the</strong> meander<strong>in</strong>g river has become braided and a new vegetation community<br />

c has arrived. The biodiversity (t3) of fl oodpla<strong>in</strong> vegetation communities has now changed (from Hughes et al., 2005)<br />

and biological parameters that are representative of ecosystem<br />

function. Whereas <strong>the</strong> fi rst is tied to <strong>the</strong> evaluation<br />

of a restoration project, <strong>the</strong> second is descriptive of <strong>the</strong><br />

state of an ecosystem. However, <strong>the</strong>y are <strong>in</strong>terrelated <strong>in</strong><br />

functional terms and decisions on what to monitor to<br />

measure ecological quality will have a signifi cant impact<br />

on our ability to evaluate ecological success <strong>in</strong> river restoration,<br />

s<strong>in</strong>ce this will depend on pre- and post-project<br />

measurements of ecosystem-relevant parameters. In<br />

Europe, <strong>the</strong> Water Framework Directive (WFD) (2000/60/<br />

EC) provides a legislative framework for water policy and<br />

among o<strong>the</strong>r th<strong>in</strong>gs aims to ma<strong>in</strong>ta<strong>in</strong> and improve <strong>the</strong><br />

aquatic environment through attention to both water<br />

quality and quantity. It requires that member states ensure<br />

good ‘chemical status’ and ‘ecological status’ of surface<br />

waters. Groundwaters must meet ‘good groundwater<br />

status’. There is no objective for riparian and fl oodpla<strong>in</strong><br />

ecosystems; however, <strong>the</strong> health of groundwaterdependent<br />

wetlands is an <strong>in</strong>dicator of ‘good groundwater<br />

status’. ‘Ecological status’ is described as ‘an expression<br />

of <strong>the</strong> quality of <strong>the</strong> structure and function<strong>in</strong>g of aquatic<br />

ecosystems associated with surface waters’ and member<br />

states are required to monitor this status. To do this <strong>the</strong>re<br />

has to be clear understand<strong>in</strong>g of what ‘ecological status’<br />

means, and much debate about this has followed publication<br />

of <strong>the</strong> Water Framework Directive (WFD) both at<br />

European Union and national levels.<br />

The ma<strong>in</strong> descriptors of ecological quality for rivers are<br />

grouped <strong>in</strong>to biological, hydrogeomorphological and<br />

physico-chemical elements (Table 6.3). Soon after adoption<br />

of <strong>the</strong> WFD (22nd December, 2000), <strong>the</strong> fi rst volume<br />

of Common Implementation Strategy (CIS) guidance was<br />

produced. Its ma<strong>in</strong> aim is to help member states share<br />

expertise on implementation of <strong>the</strong> technical aspects of <strong>the</strong><br />

WFD, <strong>in</strong>clud<strong>in</strong>g achievement of good ecological status for<br />

surface waters. Annex V of <strong>the</strong> WFD <strong>in</strong>cludes normative<br />

defi nitions of <strong>the</strong> three ecological status categories (High,<br />

Good and Moderate). The levels at which <strong>the</strong>se three categories<br />

will be set is <strong>the</strong> subject of a major EU <strong>in</strong>tercalibration<br />

process that should have been completed <strong>in</strong>


2006, but which rema<strong>in</strong>s contentious and partial. This<br />

process is be<strong>in</strong>g carried out by <strong>the</strong> ‘European Centre for<br />

Ecological Water Quality and Intercalibration’ (EEWAI),<br />

which aims to compare <strong>the</strong> different national classifi cation<br />

systems for ecological status assessment. Two key parts of<br />

<strong>the</strong> process are identifi cation of reference conditions and<br />

establishment of monitor<strong>in</strong>g protocols. It is already known<br />

that river hydromorphology will have to be restored<br />

(unless <strong>the</strong> river stretch is designated ‘heavily modifi ed’,<br />

i.e. restoration would adversely affect its function for<br />

navigation, fl ood protection or power generation) if it is<br />

prevent<strong>in</strong>g biological elements from achiev<strong>in</strong>g ‘good ecological<br />

status’ (Withr<strong>in</strong>gton, Natural England, personal<br />

communication).<br />

Individual member states have to <strong>in</strong>tegrate implementation<br />

of <strong>the</strong> WFD with <strong>the</strong>ir already established protocols<br />

for monitor<strong>in</strong>g and evaluat<strong>in</strong>g <strong>the</strong> status of <strong>the</strong>ir environment,<br />

species and habitats. In <strong>the</strong> United K<strong>in</strong>gdom, <strong>the</strong><br />

statutory conservation agencies have recently <strong>in</strong>troduced<br />

a system of ‘Common Standards Monitor<strong>in</strong>g’ (CSM) for<br />

sites designated under national legislation and European<br />

directives. Site-based conservation is a signifi cant part of<br />

biodiversity and earth science conservation <strong>in</strong> <strong>the</strong> United<br />

K<strong>in</strong>gdom and evaluation of <strong>the</strong> effectiveness of measures<br />

put <strong>in</strong>to place to achieve biodiversity conservation is <strong>the</strong><br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 95<br />

Table 6.3 Quality elements for <strong>the</strong> classifi cation of Ecological Status-<strong>River</strong>s (EU Water Framework Directive, 2000/60/EC,<br />

Annex V)<br />

Quality Element Characteristics<br />

Biological elements • Composition and abundance of aquatic fl ora<br />

• Composition and abundance of benthic <strong>in</strong>vertebrate fauna<br />

• Composition, abundance and age structure of fi sh fauna<br />

Hydrogeomorphological elements support<strong>in</strong>g<br />

<strong>the</strong> biological elements<br />

Chemical and physico-chemical elements<br />

support<strong>in</strong>g <strong>the</strong> biological elements<br />

• Hydrological regime<br />

– quantity and dynamics of water fl ow<br />

– connection to groundwater bodies<br />

• <strong>River</strong> cont<strong>in</strong>uity<br />

• Morphological conditions<br />

– river depth and width variation<br />

– structure and substrate of <strong>the</strong> river bed<br />

– structure of <strong>the</strong> riparian zone<br />

• General<br />

– <strong>the</strong>rmal conditions<br />

– oxygenation conditions<br />

– sal<strong>in</strong>ity<br />

– acidifi cation status<br />

– nutrient conditions<br />

• Specifi c pollutants<br />

– pollution by all priority substances identifi ed as be<strong>in</strong>g<br />

discharged <strong>in</strong>to <strong>the</strong> body of water<br />

– pollution by o<strong>the</strong>r substances identifi ed as be<strong>in</strong>g discharged <strong>in</strong><br />

signifi cant quantities <strong>in</strong>to <strong>the</strong> body of water<br />

ma<strong>in</strong> aim of <strong>the</strong> CSM process. The CSM guidance for<br />

monitor<strong>in</strong>g rivers designated as important by national legislation<br />

(Sites of Special Scientifi c Interest (SSSIs) under<br />

<strong>the</strong> Wildlife and Countryside Act, 1981) recommends <strong>the</strong><br />

use of fl uvial geomorphological audit to defi ne modifi cations<br />

to rivers and identify options for river restoration<br />

(JNCC, 2004). Riparian zones are <strong>in</strong>cluded <strong>in</strong> <strong>the</strong> CSM<br />

guidance for rivers, both <strong>in</strong> <strong>the</strong>ir own right as woodland,<br />

grassland or swamp communities and for <strong>the</strong> contribution<br />

<strong>the</strong>y make to <strong>in</strong>-channel river communities. Flow levels<br />

and patterns are mostly assessed <strong>in</strong> terms of <strong>the</strong>ir importance<br />

to <strong>in</strong>stream species and habitats ra<strong>the</strong>r than to adjacent<br />

terrestrial ecosystems. The functional attributes of<br />

riparian zones is considered through measurement of<br />

factors such as production of woody debris and leaf litter<br />

as is <strong>the</strong>ir <strong>in</strong>tr<strong>in</strong>sic habitat value, for example bird, mammal<br />

or <strong>in</strong>vertebrate habitat. Such monitor<strong>in</strong>g schemes for designated<br />

sites will be <strong>in</strong>tegrated with<strong>in</strong> broader-scale, river<br />

bas<strong>in</strong> wide monitor<strong>in</strong>g of ecological quality, under <strong>the</strong><br />

WFD, of both aquatic and groundwater-dependent terrestrial<br />

ecosystems.<br />

The design and density of monitor<strong>in</strong>g networks eventually<br />

established on European rivers will determ<strong>in</strong>e<br />

how useful <strong>the</strong>y are for also evaluat<strong>in</strong>g river restoration<br />

success. Numerous research projects funded by <strong>the</strong>


96 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

European Commission are <strong>in</strong> <strong>the</strong> process of explor<strong>in</strong>g<br />

monitor<strong>in</strong>g protocols for particular groups of species, such<br />

as macro<strong>in</strong>vertebrates and fi sh, with a particular emphasis<br />

on <strong>the</strong>ir transferability, comparability and ability to <strong>in</strong>dicate<br />

ecological quality. For example, <strong>the</strong> LIFE <strong>in</strong> UK<br />

<strong>River</strong>s Project (1998–2003) produced monitor<strong>in</strong>g protocols<br />

for rivers with Callitricho-Batrachion vegetation and<br />

for 10 river-dependent species such as <strong>the</strong> otter (Lutra<br />

lutra). O<strong>the</strong>rs are concerned with defi nition of reference<br />

conditions, tak<strong>in</strong>g <strong>in</strong>to account climate change (e.g.<br />

EUROLIMPACTS).<br />

6.5.3 How Do We Value Biodiversity?<br />

Most commonly, use of <strong>the</strong> term biodiversity <strong>in</strong>dicates a<br />

measure of species numbers, although habitat diversity<br />

and genetic diversity are also vitally important (Heywood,<br />

1995; Gopal and Junk, 2001). It is relatively straightforward<br />

to measure species diversity compared with ecological<br />

and genetic diversity, although <strong>the</strong> signifi cance of<br />

chang<strong>in</strong>g biodiversity rema<strong>in</strong>s elusive <strong>in</strong> all cases. In<br />

riparian and fl oodpla<strong>in</strong> habitats, levels of species diversity<br />

are related to both <strong>the</strong> ecotonal situation of <strong>the</strong>se habitats<br />

and to <strong>the</strong>ir physical heterogeneity. Lateral connectivity<br />

between riparian habitats and channels and longitud<strong>in</strong>al<br />

connectivities between upper and lower parts of river<br />

bas<strong>in</strong>s all contribute to high levels of biodiversity (Tabacchi<br />

et al., 1996). Less well understood are <strong>the</strong> vertical<br />

connectivities between groundwater zones and riparian<br />

zones but <strong>the</strong>y contribute <strong>in</strong> complex ways to habitat<br />

quality <strong>in</strong> <strong>in</strong>terstices of fl oodpla<strong>in</strong> sediments (Lambs,<br />

2004) and to diversity of organisms <strong>in</strong> <strong>the</strong> hyporheic zone<br />

(Gibert et al., 1997; Hancock et al., 2005).<br />

Species composition on fl oodpla<strong>in</strong>s has changed over<br />

long and short time scales as shown by chang<strong>in</strong>g pollen<br />

profi les from peat deposits <strong>in</strong> fl oodpla<strong>in</strong> zones (Godw<strong>in</strong>,<br />

1941), respond<strong>in</strong>g to climate change and <strong>the</strong> progression<br />

of species through <strong>the</strong> landscape. In <strong>the</strong> last few centuries,<br />

many new or alien species have arrived on fl oodpla<strong>in</strong>s as<br />

a result of <strong>in</strong>troductions and garden escapes. Because<br />

riparian and fl oodpla<strong>in</strong> zones are physically highly<br />

dynamic, and are populated by species able to cope with<br />

habitat mobility, <strong>the</strong>y provide excellent habitats for ecological<br />

pioneers, which many <strong>in</strong>vasive species can be classifi<br />

ed as (Planty-Tabacchi et al., 1996). These species have<br />

often out-competed native species because hydrological<br />

and geomorphological factors have changed and <strong>the</strong><br />

species that have evolved <strong>the</strong>ir life cycles to fi t <strong>in</strong> with<br />

natural hydrological cycles are no longer favoured. In river<br />

restoration projects today, <strong>the</strong> presence of alien species<br />

can be viewed <strong>in</strong> several ways. They can ei<strong>the</strong>r be eradicated<br />

by active processes, such as clearance, or <strong>the</strong>y can<br />

have conditions for <strong>the</strong>ir regeneration made diffi cult by<br />

alter<strong>in</strong>g physical <strong>in</strong>puts, such as fl ood tim<strong>in</strong>g to reduce<br />

<strong>the</strong>ir competitive edge. On <strong>the</strong> o<strong>the</strong>r hand, <strong>the</strong> number of<br />

alien species present <strong>in</strong> many river systems makes <strong>the</strong>ir<br />

complete eradication impossible, and acceptance of some<br />

of <strong>the</strong>m as components of <strong>the</strong> ecosystems found <strong>in</strong> riparian<br />

zones is ano<strong>the</strong>r approach. It is easier to justify this<br />

approach <strong>in</strong> view of predicted climate change, which will<br />

signifi cantly alter <strong>the</strong> range that many species can occupy<br />

(Jensen, 2004) and <strong>the</strong> hydrological patterns <strong>in</strong> river bas<strong>in</strong>s<br />

(Montgomery and Boulton, 2003) but <strong>in</strong> ways that are<br />

currently uncerta<strong>in</strong>. It is certa<strong>in</strong> that species assemblages<br />

<strong>in</strong> <strong>the</strong> landscape will change and perhaps have no presentday<br />

analogues and that an open-m<strong>in</strong>ded approach to<br />

acceptable community composition will have to be taken<br />

(Hughes et al., 2005).<br />

6.6 INTERCONNECTIVITY AND BOUNDARY<br />

CROSSING: INSTITUTIONAL COMPLEXITIES<br />

OF RESTORING FLOODPLAINS<br />

The task of restor<strong>in</strong>g fl oodpla<strong>in</strong>s poses multi-dimensional<br />

challenges to policy makers and project managers alike.<br />

Involv<strong>in</strong>g essentially a reconfi guration of <strong>the</strong> <strong>in</strong>teraction<br />

between a river and adjacent low-ly<strong>in</strong>g land, fl oodpla<strong>in</strong><br />

restoration has far-reach<strong>in</strong>g implications for exist<strong>in</strong>g<br />

forms of water and land use. Floodpla<strong>in</strong>s provide multiple<br />

functions and services for humans as well as <strong>the</strong> natural<br />

environment. These can range from valuable artefacts for<br />

socio-economic reproduction, such as crops, timber, water<br />

or prime land for development, to less tangible but equally<br />

valuable functions, such as protection from fl ood<strong>in</strong>g,<br />

attractive landscapes or opportunities for recreational pursuits.<br />

The restoration of functional fl oodpla<strong>in</strong>s requires<br />

changes to exist<strong>in</strong>g activities on <strong>the</strong> site of <strong>the</strong> fl oodpla<strong>in</strong><br />

itself, but also – particularly <strong>in</strong> <strong>the</strong> case of larger schemes<br />

– along whole reaches of a river and even a whole catchment.<br />

On this wider scale it can signifi cantly <strong>in</strong>fl uence, for<br />

<strong>in</strong>stance, levels of fl ood protection, <strong>the</strong> navigability of a<br />

river reach or <strong>the</strong> viability of current farm<strong>in</strong>g practices. In<br />

this way, fl oodpla<strong>in</strong> restoration affects a wide range of<br />

stakeholders and <strong>in</strong>terests (Adams and Perrow, 1999;<br />

Adams et al., 2004; Turner et al., 2000; Adger and Luttrell,<br />

2000), mak<strong>in</strong>g it potentially highly controversial.<br />

Beh<strong>in</strong>d <strong>the</strong>se stakeholders and <strong>in</strong>terests lie <strong>in</strong>stitutions<br />

– understood here as rule systems – designed to<br />

protect and provide a variety of private and public goods,<br />

rang<strong>in</strong>g from commercial products to rights of access. For<br />

each of <strong>the</strong> policy fi elds affected by fl oodpla<strong>in</strong> restoration<br />

– primarily water protection, fl ood defence, nature conservation,<br />

recreation, navigation, urban and rural development<br />

– complex <strong>in</strong>stitutional arrangements have been


designed and adapted over <strong>the</strong> years. Each <strong>in</strong>stitutional<br />

arrangement comprises a set of codifi ed norms (such as<br />

laws, regulations and contractual obligations), plann<strong>in</strong>g<br />

<strong>in</strong>struments and fund<strong>in</strong>g mechanisms, as well as standardised<br />

procedures of operation, values and accepted<br />

practices of <strong>the</strong> relevant organised and <strong>in</strong>dividual actors.<br />

A scheme to restore a fl oodpla<strong>in</strong> requires <strong>the</strong> successful<br />

enrolment of <strong>the</strong>se <strong>in</strong>stitutions and organisations <strong>in</strong> such<br />

a way as to create a result acceptable to <strong>the</strong> pr<strong>in</strong>cipal<br />

stakeholders. This is a highly complex process, <strong>in</strong>volv<strong>in</strong>g<br />

multiple uncerta<strong>in</strong>ties (on <strong>in</strong>stitutional constra<strong>in</strong>ts, see<br />

WWF, 2000).<br />

<strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>in</strong>terdependence of multiple functions,<br />

actors and <strong>in</strong>stitutions is, however, not <strong>the</strong> only major<br />

socio-political challenge of fl oodpla<strong>in</strong> restoration. Considerable<br />

uncerta<strong>in</strong>ty is generated by <strong>the</strong> diverse spatial scales<br />

and time frames <strong>in</strong>volved. As described above, restor<strong>in</strong>g a<br />

fl oodpla<strong>in</strong> requires consideration of <strong>the</strong> longitud<strong>in</strong>al connectivity<br />

of a fl oodpla<strong>in</strong> to river uses both up- and downstream<br />

as well as of <strong>the</strong> lateral connectivity to ways <strong>in</strong><br />

which adjacent land is used (Adams and Perrow, 1999).<br />

Even on site, <strong>in</strong>terventions generally cut across several<br />

functional and adm<strong>in</strong>istrative boundaries. These can relate<br />

to <strong>the</strong> spatial remit of local landowners and farmers, plann<strong>in</strong>g<br />

authorities, government agencies, protected areas or<br />

<strong>in</strong>frastructure networks (e.g. rail, roads, canals). Temporally,<br />

functional fl oodpla<strong>in</strong>s are characterised by <strong>the</strong>ir<br />

dependence on fl ood<strong>in</strong>g events which are, by <strong>the</strong>ir nature,<br />

periodic and unpredictable, and by signifi cant time lags<br />

between changes <strong>in</strong> biotic and abiotic systems (see above;<br />

Adams and Perrow, 1999). Socio-economically, too, <strong>the</strong><br />

process of restor<strong>in</strong>g a fl oodpla<strong>in</strong> is framed by diverse time<br />

scales, rang<strong>in</strong>g from <strong>the</strong> payback periods for <strong>in</strong>vestments<br />

<strong>in</strong> altered practices of agriculture and forestry to <strong>the</strong> electoral<br />

periods of key public authorities. Floodpla<strong>in</strong> restoration<br />

is, <strong>the</strong>refore, not only highly complex but also highly<br />

unpredictable. The <strong>in</strong>stitutional challenge is fur<strong>the</strong>r complicated<br />

by <strong>the</strong> fact that many <strong>in</strong>stitutional arrangements<br />

– <strong>in</strong> particular for nature conservation – exhibit a strong<br />

tendency to protect exist<strong>in</strong>g conditions ra<strong>the</strong>r than encourage<br />

change, and those which do pursue change are generally<br />

geared towards achiev<strong>in</strong>g specifi c targets ra<strong>the</strong>r than<br />

creat<strong>in</strong>g suitable frameworks for open-ended processes, as<br />

is required for functional fl oodpla<strong>in</strong>s.<br />

6.6.1 Effective Institutions: The Search for Optimal<br />

Fit, Interplay and Scale<br />

Our knowledge of <strong>in</strong>stitutions which can support – or<br />

obstruct – <strong>the</strong> protection of public goods such as water,<br />

fl ood defence and biodiversity has been develop<strong>in</strong>g rapidly<br />

over <strong>the</strong> past decade (cf. Breit et al., 2003). However, rela-<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 97<br />

tively little is known about <strong>the</strong> <strong>in</strong>stitutional dimensions of<br />

fl oodpla<strong>in</strong> restoration itself. Exceptions <strong>in</strong>clude studies<br />

of <strong>in</strong>stitutional constra<strong>in</strong>ts and complexities (Adams and<br />

Perrow, 1999), compet<strong>in</strong>g discourses of fl oodpla<strong>in</strong> restoration<br />

(Adams et al., 2004), relevant European Union policies<br />

(WWF, 2000, 2004) and European case studies (e.g. Zöckler,<br />

2000) and economic valuations of <strong>the</strong> functions and services<br />

provided by fl oodpla<strong>in</strong>s or wetlands (Gren et al.,<br />

1995; Turner et al., 2000; Adger and Luttrell, 2000).<br />

The Science Plan of <strong>the</strong> Institutional Dimensions of<br />

Global Environmental Change Project (IDGEC) of <strong>the</strong><br />

International Human Dimensions Programme (IHDP)<br />

offers useful analytical frameworks for conceptualis<strong>in</strong>g<br />

some of <strong>the</strong> essential <strong>in</strong>stitutional challenges of resource<br />

management <strong>in</strong> general and fl oodpla<strong>in</strong> restoration <strong>in</strong> particular<br />

(Young, 1999, 2002). It identifi es three generic<br />

factors <strong>in</strong>fl uenc<strong>in</strong>g <strong>the</strong> effectiveness of environmental<br />

<strong>in</strong>stitutions: problems of fi t, problems of <strong>in</strong>terplay and<br />

problems of scale.<br />

The issue of fi t addresses <strong>the</strong> need to develop <strong>in</strong>stitutional<br />

arrangements that match <strong>the</strong> properties of <strong>the</strong> biogeophysical<br />

systems <strong>the</strong>y are designed to regulate. Fit can<br />

relate to a variety of ecosystem properties. The follow<strong>in</strong>g<br />

are identifi ed <strong>in</strong> <strong>the</strong> IDGEC Science Plan: closed vs open<br />

systems; heterogeneity/homogeneity; <strong>in</strong>terdependencies<br />

among subsystems; simplicity/complexity; productivity/<br />

metabolism; cyclicity/periodicity; resilience; equilibria;<br />

dynamics (Young, 1999, p. 47). Problems of spatial misfi t<br />

are a particularly common cause of <strong>in</strong>stitutional <strong>in</strong>effectiveness.<br />

The territories covered by <strong>in</strong>stitutions rarely<br />

match those of biogeophysical systems, result<strong>in</strong>g <strong>in</strong> an<br />

<strong>in</strong>ability of <strong>the</strong> <strong>in</strong>stitutions to <strong>in</strong>ternalise external effects<br />

(both positive and negative) effectively. The management<br />

of fl oodpla<strong>in</strong>s is fraught with boundary problems of this<br />

k<strong>in</strong>d. Floodpla<strong>in</strong> restoration not only works across a<br />

variety of physical spaces along <strong>the</strong> river and across <strong>the</strong><br />

catchment, but also <strong>in</strong>volves <strong>in</strong>stitutions and organisations<br />

from multiple policy fi elds – from nature conservation and<br />

fl ood defence to agriculture – each with <strong>the</strong>ir own spatial<br />

remits.<br />

Interplay relates, by contrast, to <strong>in</strong>terdependencies<br />

between different <strong>in</strong>stitutions. The assumption here is that<br />

<strong>the</strong> effectiveness of an <strong>in</strong>stitution depends not only on its<br />

<strong>in</strong>herent qualities but also on how well it builds on, and is<br />

connected to, <strong>the</strong> broader <strong>in</strong>stitutional context. Institutional<br />

<strong>in</strong>terplay can be horizontal, between different policy<br />

fi elds, and vertical, between different levels of social<br />

organisation. A fur<strong>the</strong>r dist<strong>in</strong>ction is made between functional<br />

l<strong>in</strong>kages emanat<strong>in</strong>g from <strong>the</strong> properties of <strong>the</strong> <strong>in</strong>stitutions<br />

<strong>in</strong>volved and political l<strong>in</strong>kages as expressions of<br />

deliberation (Young, 2002). Problems of <strong>in</strong>terplay are very<br />

familiar to efforts to restore fl oodpla<strong>in</strong>s, which are often


98 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

confounded by <strong>the</strong> <strong>in</strong>ability to bridge differences <strong>in</strong> <strong>the</strong><br />

objectives, power structures and modes of action of <strong>the</strong><br />

various key organisations. Horizontal <strong>in</strong>terplay is complicated<br />

by <strong>the</strong> number of policy fi elds affected and vertical<br />

<strong>in</strong>terplay by <strong>the</strong> <strong>in</strong>creas<strong>in</strong>g role of <strong>the</strong> European Union and<br />

catchment-scale approaches to fl oodpla<strong>in</strong> management.<br />

Problems of scale can be of a spatial and a temporal<br />

nature. Spatially <strong>the</strong> effectiveness of an <strong>in</strong>stitution depends<br />

on fi nd<strong>in</strong>g <strong>the</strong> appropriate level of social organisation for<br />

specifi c <strong>in</strong>struments and measures, tak<strong>in</strong>g consideration of<br />

<strong>the</strong> connectivity between scales and <strong>the</strong> needs this creates<br />

for multi-level and multi-directional forms of governance.<br />

Temporally <strong>the</strong> issue of scale is about manag<strong>in</strong>g <strong>the</strong><br />

diverse time frames with<strong>in</strong> which actors operate, <strong>in</strong>stitutions<br />

work, projects are implemented, ideas are generated<br />

etc. Here, too, <strong>the</strong> relevance to fl oodpla<strong>in</strong> restoration is<br />

self-evident. Identify<strong>in</strong>g <strong>the</strong> appropriate spatial scale for<br />

policy development, strategic guidance, operational management,<br />

public participation and so on is of paramount<br />

importance. Similarly, actors operate accord<strong>in</strong>g to very<br />

different time scales, some of which are rigid and predictable,<br />

o<strong>the</strong>rs much less so.<br />

Research on problems of fi t, <strong>in</strong>terplay and scale suggests<br />

that solutions are rarely straightforward. For <strong>in</strong>stance,<br />

efforts to overcome problems of spatial fi t by <strong>in</strong>stitutionalis<strong>in</strong>g<br />

river bas<strong>in</strong> management can create new misfi ts and<br />

disturb exist<strong>in</strong>g modes of <strong>in</strong>terplay (Moss, 2003). Success<br />

would appear to be dependent less on attempt<strong>in</strong>g to reduce<br />

<strong>the</strong> given complexities and more on fi nd<strong>in</strong>g ways of<br />

accommodat<strong>in</strong>g complexity. It is argued here that a similar<br />

approach is required when deal<strong>in</strong>g with uncerta<strong>in</strong>ty. Given<br />

that substantial uncerta<strong>in</strong>ties will cont<strong>in</strong>ue to surround<br />

fl oodpla<strong>in</strong> restoration despite advances <strong>in</strong> our knowledge<br />

of <strong>the</strong> physical, biological and socio-political systems, it<br />

makes sense to consider possible cop<strong>in</strong>g strategies. The<br />

rema<strong>in</strong>der of this section addresses ways of cop<strong>in</strong>g with<br />

<strong>in</strong>stitutional uncerta<strong>in</strong>ty <strong>in</strong> fl oodpla<strong>in</strong> restoration with an<br />

empirically based study of three different approaches.<br />

6.6.2 Cop<strong>in</strong>g with <strong>Uncerta<strong>in</strong>ty</strong> I:<br />

Keep<strong>in</strong>g <strong>Restoration</strong> Simple<br />

The fi rst option for cop<strong>in</strong>g with uncerta<strong>in</strong>ty is to limit <strong>the</strong><br />

scope and scale of restoration. This was particularly<br />

common of <strong>the</strong> earlier schemes to restore fl oodpla<strong>in</strong>s <strong>in</strong><br />

Europe. Up until <strong>the</strong> late 1990s most fl oodpla<strong>in</strong> restoration<br />

schemes were small-scale and site-based. They were<br />

typically s<strong>in</strong>gle-issue projects, target<strong>in</strong>g environmental<br />

improvements as a rule. They <strong>in</strong>volved only a small<br />

number of actors and policy <strong>in</strong>struments, often rely<strong>in</strong>g on<br />

a s<strong>in</strong>gle source of fund<strong>in</strong>g for <strong>the</strong> physical <strong>in</strong>terventions.<br />

It is generally true to say that <strong>the</strong>se early generation res-<br />

toration schemes were conducted largely <strong>in</strong> isolation from<br />

national or regional policy <strong>in</strong>itiatives, whe<strong>the</strong>r for fl ood<br />

protection, biodiversity enhancement or rural development.<br />

Examples <strong>in</strong>clude <strong>the</strong> Rhe<strong>in</strong>vorland-Süd project <strong>in</strong><br />

Germany, a scheme to improve hydrological and ecological<br />

conditions by widen<strong>in</strong>g ducts and remov<strong>in</strong>g structures<br />

<strong>in</strong> a section of <strong>the</strong> Rh<strong>in</strong>e fl oodpla<strong>in</strong> near Rastatt, <strong>the</strong> Long<br />

Eau project <strong>in</strong> England, <strong>in</strong> which fl ood banks were set back<br />

for primarily environmental benefi ts, and <strong>the</strong> Bourret<br />

project <strong>in</strong> France, a scheme to reconnect an old arm to <strong>the</strong><br />

<strong>River</strong> Garonne and restore an alluvial forest – aga<strong>in</strong> primarily<br />

for environmental benefi ts.<br />

Be<strong>in</strong>g relatively unambitious and straightforward,<br />

schemes of this k<strong>in</strong>d tend to avoid <strong>the</strong> most press<strong>in</strong>g problems<br />

associated with high levels of uncerta<strong>in</strong>ty and complexity.<br />

They have succeeded <strong>in</strong> restor<strong>in</strong>g fl oodpla<strong>in</strong><br />

habitats with limited resources and, <strong>in</strong> some cases, with<strong>in</strong><br />

a short period. With <strong>the</strong> benefi t of relatively straightforward<br />

adm<strong>in</strong>istrative procedures, organisational structures<br />

and fund<strong>in</strong>g mechanisms it has been demonstrated how<br />

fl oodpla<strong>in</strong> ecosystems can be restored on a small scale.<br />

These schemes do, however, have several critical limitations.<br />

They rarely <strong>in</strong>corporate a catchment perspective on<br />

restoration, but concentrate on <strong>the</strong> site itself. Be<strong>in</strong>g predom<strong>in</strong>antly<br />

s<strong>in</strong>gle-issue schemes, <strong>the</strong>y regularly overlook<br />

potential benefi ts for o<strong>the</strong>r policy areas, such as fl ood<br />

protection, recreation or rural development. Little attention<br />

is generally paid to cultivat<strong>in</strong>g support for <strong>the</strong> project<br />

<strong>in</strong> <strong>the</strong> wider policy mak<strong>in</strong>g doma<strong>in</strong>, scientifi c communities<br />

or even <strong>in</strong> <strong>the</strong> local community. The performance of<br />

many such projects is rarely monitored or evaluated<br />

systematically.<br />

In terms of resolv<strong>in</strong>g problems of fi t, <strong>in</strong>terplay and scale<br />

it is possible to observe how schemes of this k<strong>in</strong>d are illequipped<br />

to meet <strong>the</strong> pr<strong>in</strong>cipal <strong>in</strong>stitutional challenges to<br />

fl oodpla<strong>in</strong> restoration set out above. Spatially, <strong>the</strong> small<br />

scale and site focus of <strong>the</strong> projects offer little opportunity<br />

to consider <strong>the</strong> catchment dimensions of fl ow regimes and<br />

biophysical connectivity beyond <strong>the</strong> immediate reach.<br />

Institutional <strong>in</strong>terplay may be less diffi cult but only<br />

because <strong>the</strong> number of policy fi elds and organisations<br />

<strong>in</strong>volved is kept small. Integrat<strong>in</strong>g <strong>the</strong> schemes <strong>in</strong>to <strong>the</strong><br />

development of <strong>the</strong> locality or region may well prove diffi<br />

cult at a later date for this reason. Problems of scale are<br />

reduced to site-based perspectives, miss<strong>in</strong>g valuable<br />

opportunities to explore multi-scalar solutions.<br />

6.6.3 Cop<strong>in</strong>g with <strong>Uncerta<strong>in</strong>ty</strong> II: Embrac<strong>in</strong>g <strong>the</strong><br />

Challenge of Open Outcomes<br />

Follow<strong>in</strong>g recent shifts <strong>in</strong> policies towards fl ood risk management,<br />

<strong>in</strong>tegrated water resources management and


ural development, more favourable framework conditions<br />

are creat<strong>in</strong>g w<strong>in</strong>dows of opportunity for ambitious forms<br />

of fl oodpla<strong>in</strong> restoration (see previously). In response to<br />

<strong>the</strong>se <strong>in</strong>stitutional drivers, and also to a grow<strong>in</strong>g recognition<br />

of <strong>the</strong> <strong>in</strong>adequacies of current ways of manag<strong>in</strong>g<br />

rivers, a new generation of fl oodpla<strong>in</strong> restoration schemes<br />

is emerg<strong>in</strong>g which are of a quite different scope to those<br />

of <strong>the</strong> early to mid-1990s. These schemes deliberately set<br />

out to address some of <strong>the</strong> complex challenges to largescale,<br />

<strong>in</strong>tegrated fl oodpla<strong>in</strong> restoration. Dist<strong>in</strong>ctive features<br />

of <strong>the</strong> new generation schemes are <strong>the</strong>ir multiple<br />

objectives (cover<strong>in</strong>g, for <strong>in</strong>stance, fl ood defence, biodiversity,<br />

rural development and water quality management),<br />

<strong>the</strong>ir wide actor engagement (<strong>in</strong>clud<strong>in</strong>g <strong>the</strong> relevant policy<br />

fi elds, local authorities, non-government organisations<br />

and <strong>the</strong> general public), <strong>the</strong>ir use of various <strong>in</strong>struments<br />

from different policy fi elds (e.g. jo<strong>in</strong>t fund<strong>in</strong>g from fl ood<br />

defence and agri-environment budgets) and <strong>the</strong>ir <strong>in</strong>teraction<br />

with policy mak<strong>in</strong>g processes, serv<strong>in</strong>g for <strong>in</strong>stance as<br />

pilot projects for national policy development. In addition,<br />

<strong>the</strong>y often have a long term vision for <strong>the</strong> measures envisaged<br />

and take a catchment – or at least large-scale – perspective<br />

on <strong>the</strong> fl oodpla<strong>in</strong>. <strong>Restoration</strong> sites are selected<br />

accord<strong>in</strong>g to <strong>the</strong>ir suitability for <strong>the</strong> catchment and not<br />

primarily because <strong>the</strong>y are available.<br />

Examples <strong>in</strong>clude: <strong>the</strong> Lenzen project <strong>in</strong> Germany, a<br />

major scheme to set back fl ood banks along a length of<br />

<strong>the</strong> Elbe river allow<strong>in</strong>g fl ood<strong>in</strong>g primarily for nature conservation,<br />

but also fl ood defence and regional development,<br />

benefi ts; <strong>the</strong> Parrett Catchment Project <strong>in</strong> England,<br />

an ongo<strong>in</strong>g project to promote more susta<strong>in</strong>able techniques<br />

of fl ood management <strong>in</strong> <strong>the</strong> whole catchment<br />

serv<strong>in</strong>g multiple purposes (fl ood protection, water level<br />

regulation, biodiversity targets, rural development); and<br />

<strong>the</strong> La Bassée project <strong>in</strong> France, a planned, large-scale<br />

fl ood retention scheme on <strong>the</strong> Se<strong>in</strong>e upstream of Paris with<br />

multiple benefi ts (fl ood protection, biodiversity, regional<br />

development). S<strong>in</strong>ce <strong>the</strong>se new generation schemes were<br />

only launched from <strong>the</strong> late 1990s onwards and are all at<br />

very early stages of implementation it is at present impossible<br />

to judge <strong>the</strong>ir effectiveness. They would at least<br />

appear to have <strong>the</strong> potential to overcome some of <strong>the</strong><br />

pr<strong>in</strong>cipal <strong>in</strong>stitutional constra<strong>in</strong>ts to fl oodpla<strong>in</strong> restoration<br />

which have thwarted or curtailed efforts <strong>in</strong> <strong>the</strong> past.<br />

Our research fi nd<strong>in</strong>gs, however, caution aga<strong>in</strong>st overoptimistic<br />

expectations from <strong>the</strong> new generation of projects.<br />

Early signs suggest that <strong>the</strong> sheer complexity of <strong>the</strong><br />

tasks <strong>the</strong>y are tackl<strong>in</strong>g and <strong>the</strong> uncerta<strong>in</strong>ties <strong>the</strong>y are<br />

expos<strong>in</strong>g are pos<strong>in</strong>g a major problem for project management.<br />

Build<strong>in</strong>g and ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g <strong>the</strong> large partnerships<br />

takes time and care. Strik<strong>in</strong>g an acceptable balance and<br />

negotiat<strong>in</strong>g trade-offs between diverse policy objectives<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 99<br />

is very demand<strong>in</strong>g. Access<strong>in</strong>g multiple fund<strong>in</strong>g sources<br />

requires a high degree of fl exibility to satisfy different<br />

fund<strong>in</strong>g agencies. Attempts to enroll <strong>in</strong>struments from<br />

different policy fi elds can reveal serious <strong>in</strong>compatibilities<br />

and <strong>in</strong>consistencies. As a result, project design and<br />

implementation has become more complex, more timeconsum<strong>in</strong>g<br />

and more expensive, endanger<strong>in</strong>g effective<br />

implementation.<br />

Floodpla<strong>in</strong> restoration schemes of this k<strong>in</strong>d are mak<strong>in</strong>g<br />

substantial steps towards address<strong>in</strong>g problems of fi t, <strong>in</strong>terplay<br />

and scale. Their catchment orientation and long term<br />

perspective create a better fi t between ecosystem properties<br />

of <strong>the</strong> fl oodpla<strong>in</strong> and <strong>the</strong> <strong>in</strong>stitutional arrangements<br />

for its restoration, both <strong>in</strong> spatial and temporal terms.<br />

Build<strong>in</strong>g on better <strong>in</strong>terplay between <strong>in</strong>stitutions is central<br />

to <strong>the</strong> new generation projects, as is exploit<strong>in</strong>g different<br />

scales of action – from national pilots to local management<br />

teams – for different purposes. What <strong>the</strong> schemes<br />

are reveal<strong>in</strong>g, though, are serious secondary problems<br />

associated with this more ambitious and <strong>in</strong>tegrated<br />

approach to fl oodpla<strong>in</strong> restoration. Efforts to take on problems<br />

of fi t, <strong>in</strong>terplay and scale are prov<strong>in</strong>g, <strong>in</strong> many cases,<br />

too demand<strong>in</strong>g for project management. This does not<br />

query <strong>the</strong> desirability of address<strong>in</strong>g <strong>the</strong>se core <strong>in</strong>stitutional<br />

problems but, ra<strong>the</strong>r, raises questions about how project<br />

managers can be assisted <strong>in</strong> do<strong>in</strong>g so.<br />

6.6.4 Cop<strong>in</strong>g with <strong>Uncerta<strong>in</strong>ty</strong> III: Tighten<strong>in</strong>g<br />

Controls to Secure Better Policy Delivery<br />

A third way of cop<strong>in</strong>g with uncerta<strong>in</strong>ty can be identifi ed<br />

not at <strong>the</strong> level of <strong>in</strong>dividual schemes of fl oodpla<strong>in</strong> restoration<br />

but <strong>in</strong> <strong>the</strong> development and pursuit of policy. Here <strong>the</strong><br />

uncerta<strong>in</strong>ty addressed relates to <strong>the</strong> outcomes of new policies<br />

and <strong>the</strong> strategy is to attempt to m<strong>in</strong>imise uncerta<strong>in</strong>ties<br />

of policy delivery by tighten<strong>in</strong>g controls over those<br />

entrusted with implementation. As described earlier, <strong>in</strong><br />

several policy fi elds of direct relevance to fl oodpla<strong>in</strong><br />

restoration a more <strong>in</strong>tegrated and holistic approach to<br />

problem solv<strong>in</strong>g can be detected. This applies particularly<br />

to fl ood protection, follow<strong>in</strong>g recent fl ood<strong>in</strong>g events, water<br />

resources management, <strong>in</strong> response to <strong>the</strong> Water Framework<br />

Directive, and nature conservation. Changes <strong>in</strong><br />

policy content potentially conducive to fl oodpla<strong>in</strong> restoration<br />

can also be observed – if to a lesser degree – across<br />

Europe <strong>in</strong> spatial plann<strong>in</strong>g, agriculture, forestry and<br />

rural development. Many of <strong>the</strong>se supranational and<br />

national policy <strong>in</strong>itiatives are characterised by a more<br />

comprehensive problem analysis, longer term visions for<br />

improvements, better cross-sectoral policy <strong>in</strong>tegration,<br />

more strategic guidance and stronger and broader local<br />

partnerships. The guid<strong>in</strong>g pr<strong>in</strong>ciples underp<strong>in</strong>n<strong>in</strong>g


100 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Integrated Water Resources Management (IWRM) on a<br />

river bas<strong>in</strong> scale are a case <strong>in</strong> po<strong>in</strong>t (WWF, 2004, p. 24).<br />

Ironically, whilst policy content is, generally speak<strong>in</strong>g,<br />

tak<strong>in</strong>g a broader perspective, modes of policy implementation<br />

are to some extent becom<strong>in</strong>g more restrictive. In<br />

recent years a grow<strong>in</strong>g array of <strong>in</strong>struments have been<br />

<strong>in</strong>troduced which set out <strong>in</strong> detail not only what policy is<br />

to be pursued but also how this is to be done at <strong>the</strong> operational<br />

level. Targets are set to measure progress, strict<br />

consultation procedures must be followed to ga<strong>in</strong> plann<strong>in</strong>g<br />

approval, match-fund<strong>in</strong>g is required to demonstrate multifunctionality<br />

and audits are conducted to assess <strong>the</strong> performance<br />

of projects and programmes alike. These controls<br />

are widely justifi ed by governments <strong>in</strong> terms of <strong>the</strong>ir value<br />

<strong>in</strong> improv<strong>in</strong>g accountability, policy <strong>in</strong>tegration and, above<br />

all, cost effectiveness.<br />

Our research suggests, however, that measures of this<br />

k<strong>in</strong>d are hav<strong>in</strong>g important (un<strong>in</strong>tended) negative effects on<br />

<strong>the</strong> ability of project managers to implement fl oodpla<strong>in</strong><br />

restoration schemes. The fi rst problem relates to <strong>the</strong> cumulative<br />

effect of <strong>the</strong> new policy <strong>in</strong>itiatives. Each of <strong>the</strong> new<br />

requirements – whe<strong>the</strong>r on policy content or style – may<br />

<strong>in</strong>dividually make a lot of sense. Experiences of policy<br />

implementation show that <strong>the</strong> comb<strong>in</strong>ed effect of multiple<br />

new requirements can be to create a degree of management<br />

complexity that can severely delay <strong>the</strong> progress of<br />

some restoration projects and cause o<strong>the</strong>rs to be shelved.<br />

Ironically, <strong>the</strong>refore, effective policy delivery is be<strong>in</strong>g<br />

jeopardised by <strong>the</strong> sheer extent of policy reform.<br />

The second problem is more fundamental, hav<strong>in</strong>g to<br />

do with an emergent culture of control <strong>in</strong> policy mak<strong>in</strong>g<br />

circles. The measures to <strong>in</strong>crease accountability, policy<br />

<strong>in</strong>tegration and cost effectiveness refl ect not only very<br />

justifi able concerns about effective policy implementation<br />

and effi cient use of public funds but also <strong>the</strong> concerns of<br />

senior management <strong>in</strong> many government agencies that <strong>the</strong><br />

policy reth<strong>in</strong>k<strong>in</strong>g described above is not fi lter<strong>in</strong>g down<br />

effectively to <strong>the</strong> operational level. This argument is used<br />

by senior offi cers to justify tighter control – or ‘guidance’<br />

– to assure more effective implementation. Whilst <strong>the</strong> need<br />

for greater strategic guidance over such complex issues as<br />

a catchment-scale approach to fl ood protection is undisputed<br />

by all those <strong>in</strong>volved, one of <strong>the</strong> effects has been to<br />

restrict <strong>the</strong> freedom of action of project and programme<br />

managers. In <strong>the</strong> past <strong>the</strong>ir judgement – for example on<br />

whe<strong>the</strong>r to fund a restoration project – was based on <strong>the</strong>ir<br />

<strong>in</strong>dividual expertise, local knowledge and professional<br />

experience; today it is framed much more by targets<br />

devised at regional, national or even supranational levels.<br />

Consequently, <strong>the</strong> nature of <strong>the</strong>ir work is adapt<strong>in</strong>g <strong>in</strong> order<br />

to meet what Michael Power has termed <strong>the</strong> ‘rituals of<br />

verifi cation’ required by audit<strong>in</strong>g processes (Power, 1997).<br />

For <strong>the</strong> task of restor<strong>in</strong>g fl oodpla<strong>in</strong>s this poses a particular<br />

dilemma: whilst recent policy shifts and new generation<br />

schemes are encourag<strong>in</strong>g fl oodpla<strong>in</strong> restoration to enterta<strong>in</strong><br />

greater risks and uncerta<strong>in</strong>ties, adm<strong>in</strong>istrative procedures<br />

to assure policy implementation are becom<strong>in</strong>g<br />

<strong>in</strong>creas<strong>in</strong>gly risk-averse.<br />

In terms of fi t, <strong>in</strong>terplay and scale <strong>the</strong> picture here is<br />

more differentiated. Recent policy <strong>in</strong>itiatives relat<strong>in</strong>g to<br />

fl oodpla<strong>in</strong> restoration are certa<strong>in</strong>ly address<strong>in</strong>g very clearly<br />

problems of spatial and temporal fi t, tak<strong>in</strong>g a more catchment-oriented<br />

and long term perspective on <strong>the</strong> river and<br />

land management. Inter-sectoral <strong>in</strong>terplay is also strong.<br />

Vertical <strong>in</strong>terplay and issues of scale appear more problematic,<br />

however. The rhetoric of policy documents tend<br />

to be very supportive of multi-level and multi-direction<br />

governance. The reality – whe<strong>the</strong>r <strong>in</strong>tentional or not – is<br />

often very different, with control mechanisms of central<br />

government agencies reach<strong>in</strong>g down <strong>in</strong>to <strong>the</strong> operational<br />

level of project management to an unprecedented extent.<br />

This, it appears, is underm<strong>in</strong><strong>in</strong>g both project implementation<br />

and – ultimately – policy delivery.<br />

6.6.5 Mak<strong>in</strong>g Policy More Sensitive to <strong>the</strong> Challenges<br />

of Project Management<br />

It has been observed here how policy makers, <strong>in</strong> <strong>the</strong>ir<br />

efforts to encourage <strong>in</strong>tegrated, cross-sectoral and multiagency<br />

action, often overlook <strong>the</strong> implications for implementation<br />

at <strong>the</strong> operational level. In future, more<br />

consideration needs to be given to how <strong>in</strong>dividual policy<br />

<strong>in</strong>centives work <strong>in</strong> conjunction with o<strong>the</strong>rs; that is, how<br />

<strong>the</strong>y alter <strong>the</strong> exist<strong>in</strong>g <strong>in</strong>stitutional sett<strong>in</strong>g. In addition,<br />

policy makers need to be more sensitive to <strong>the</strong> contexts of<br />

action <strong>in</strong> which <strong>the</strong>ir <strong>in</strong>struments operate. What makes a<br />

policy <strong>in</strong>strument effective is not <strong>the</strong> assumed preferences<br />

of <strong>in</strong>dividuals act<strong>in</strong>g accord<strong>in</strong>g to a rational choice logic<br />

but <strong>the</strong> real scope and will<strong>in</strong>gness of stakeholders to alter<br />

<strong>the</strong>ir practices. More feedback <strong>in</strong>to policy mak<strong>in</strong>g processes<br />

is needed about <strong>the</strong> real-life experiences of project<br />

management and stakeholder <strong>in</strong>volvement at <strong>the</strong> operational<br />

level. This will require more monitor<strong>in</strong>g of how and<br />

why policy <strong>in</strong>struments do or do not work <strong>in</strong> practice. In<br />

addition, if uncerta<strong>in</strong>ty and complexity are unavoidable<br />

features of ambitious schemes of fl oodpla<strong>in</strong> restoration, as<br />

it appears, <strong>the</strong>n project managers and o<strong>the</strong>r stakeholders<br />

need assistance <strong>in</strong> assess<strong>in</strong>g <strong>the</strong> situation and <strong>the</strong>ir own<br />

ability to meet <strong>the</strong> challenge. Policy needs to provide not<br />

only targets for orientation but also frameworks for develop<strong>in</strong>g<br />

<strong>the</strong> necessary economic, social and <strong>in</strong>stitutional<br />

capital at local and regional level, and <strong>in</strong>struments capable<br />

of adapt<strong>in</strong>g to <strong>the</strong> dynamics of a fl oodpla<strong>in</strong> restoration<br />

process. On this basis policy makers and project managers


alike should be better equipped to identify and exploit<br />

w<strong>in</strong>dows of opportunity for <strong>the</strong> restoration of fl oodpla<strong>in</strong>s<br />

<strong>in</strong> <strong>the</strong> future.<br />

6.7 CONCLUSION<br />

A case has been made for <strong>the</strong> re-<strong>in</strong>troduction of diversity<br />

and variability <strong>in</strong> <strong>the</strong> riparian zone. It is advocated that<br />

this is done through <strong>the</strong> management of physical processes<br />

and with<strong>in</strong> a policy context that is sensitive to <strong>the</strong> complexities<br />

of successfully implement<strong>in</strong>g river restoration<br />

<strong>in</strong>itiatives. Inevitably <strong>the</strong> re-<strong>in</strong>troduction of diversity and<br />

variability will also <strong>in</strong>troduce a signifi cant level of uncerta<strong>in</strong>ty<br />

for river managers. The challenge is to decide what<br />

are acceptable levels of uncerta<strong>in</strong>ty for different stakeholders,<br />

for different scales of <strong>in</strong>volvement (such as catchment<br />

strategies or plann<strong>in</strong>g for a particular reach) and for<br />

different purposes (such as biodiversity targets or fl ood<br />

management). In <strong>the</strong> light of projected climate changes it<br />

is suggested that <strong>the</strong> key is to fi nd ways of adapt<strong>in</strong>g to<br />

uncerta<strong>in</strong>ty (see Chapter 14), ra<strong>the</strong>r than aim<strong>in</strong>g to reduce<br />

uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> scientifi c sense of <strong>the</strong> word.<br />

6.8 ACKNOWLEDGEMENTS<br />

Much of <strong>the</strong> th<strong>in</strong>k<strong>in</strong>g that has gone <strong>in</strong>to this chapter has<br />

arisen from <strong>the</strong> work of <strong>the</strong> EC-funded project FLOBAR2<br />

(EVK1-CT-1999-00031). We thank all our colleagues on<br />

that project for stimulat<strong>in</strong>g and enjoyable discussions over<br />

many years of work<strong>in</strong>g toge<strong>the</strong>r. We would also like to<br />

thank Stewart Rood, John Mahoney, David Hulse and Joan<br />

Baker for permission to use <strong>the</strong>ir work <strong>in</strong> <strong>the</strong> two case<br />

study boxes; David Withr<strong>in</strong>gton of Natural England for<br />

read<strong>in</strong>g and comment<strong>in</strong>g on this manuscript; and Ian<br />

Agnew of <strong>the</strong> Department of Geography at <strong>the</strong> University<br />

of Cambridge for draw<strong>in</strong>g <strong>the</strong> fi gures.<br />

REFERENCES<br />

Adams WM, Perrow M. 1999. Scientifi c and <strong>in</strong>stitutional constra<strong>in</strong>ts<br />

on <strong>the</strong> restoration of European fl oodpla<strong>in</strong>s. In: Marriott<br />

S, Alexander J, Hey, R (Eds), Floodpla<strong>in</strong>s: Interdiscipl<strong>in</strong>ary<br />

Approaches. Geological Society of London Special Publ. 163,<br />

Geological Society: London; 89–97.<br />

Adams WM, Perrow MR, Carpenter A. 2004. Conservatives and<br />

champions: river managers and river restoration discourses <strong>in</strong><br />

<strong>the</strong> United K<strong>in</strong>gdom. Environment and Plann<strong>in</strong>g A 36: 1929–<br />

1942.<br />

Adger WN, Luttrell C. 2000. Property rights and <strong>the</strong> utilisation<br />

of wetlands. Ecological Economics 35: 75–89.<br />

Anderson DH, Dugger BD. 1998. A conceptual basis for evaluat<strong>in</strong>g<br />

restoration success. In: Wadsworth KG (Ed.), Chang<strong>in</strong>g<br />

Resource Values <strong>in</strong> Challeng<strong>in</strong>g Times. Transactions, 63rd<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 101<br />

North American Wildlife and Natural Resources Conference,<br />

Wildlife Management Institute: Wash<strong>in</strong>gton, DC; 111–120.<br />

Arth<strong>in</strong>gton AH. 1998. Comparative Evaluation of Environmental<br />

Flow Assessment Techniques:Review of Holistic methodologies.<br />

LWRRDC Occasional Paper 26/98.<br />

Arth<strong>in</strong>gton AH, Brizga SO, Choy SC et al. 2000. Environmental<br />

Flow Requirements of <strong>the</strong> Brisbane <strong>River</strong> Downstream from<br />

Wivenhoe Dam. South East Queensland Water Corporation<br />

Ltd. and Centre for Catchment and In-stream Research, Griffi th<br />

University: Brisbane, Queensland, Australia.<br />

Baker JP, Hulse DW, Gregory SV et al. 2004. Alternative futures<br />

for <strong>the</strong> Willamette <strong>River</strong> Bas<strong>in</strong>, Oregon. Ecological Applications<br />

14 (2): 313–324.<br />

Breit H, Engels A, Moss T, Troja M. 2003. How Institutions<br />

Change. Perspectives on Social Learn<strong>in</strong>g <strong>in</strong> Global and Local<br />

Environmental Contexts. Leske and Budrich: Opladen,<br />

Germany.<br />

Bressers H, Kuks S. 2003. What does governance mean? From<br />

conception to elaboration. In: Bressers H, Rosenbaum W (Eds),<br />

Achiev<strong>in</strong>g Susta<strong>in</strong>able Development: <strong>the</strong> Challenge of Governance<br />

Across Spatial Scales. Praeger: New York.<br />

Brierley GJ, Fryirs K. 2000. <strong>River</strong> styles, a geomorphic approach<br />

to catchment characterization: implications for river rehabilitation<br />

<strong>in</strong> Bega Catchment, New South Wales, Australia. Environmental<br />

Management 25: 661–679.<br />

Brizga SO. 2000. Burnett <strong>River</strong> water allocation management<br />

plan:Proposed environmental fl ow performance measures.<br />

Department of Natural Resources: Brisbane, Australia.<br />

Brown CA, Joubert A. 2003. Us<strong>in</strong>g multicriteria analysis to<br />

develop environmental fl ow scenarios for rivers targeted for<br />

water resource management. Water South Africa 29 (4):<br />

365–378.<br />

Connell JH. 1978. Diversity <strong>in</strong> tropical ra<strong>in</strong> forests and coral<br />

reefs. Science 199: 1302–1310.<br />

Dewberry C, Burns P, Hood L. 1998. After <strong>the</strong> fl ood: <strong>the</strong> effects<br />

of <strong>the</strong> storms of 1996 on a creek restoration. <strong>Restoration</strong> and<br />

Management Notes 16: 174–182.<br />

Dodds WK, Gido K, Whiles MR et al. 2004. Life on <strong>the</strong> edge:<br />

The ecology of Great Pla<strong>in</strong>s prairie streams. BioScience 54:<br />

205–216.<br />

Dole D, Niemi E. 2004. Future water allocation and <strong>in</strong>-stream<br />

values <strong>in</strong> <strong>the</strong> Willamette <strong>River</strong> Bas<strong>in</strong>: a bas<strong>in</strong>-wide analysis.<br />

Ecological Applications 14: 355–367.<br />

Downs PW, Sk<strong>in</strong>ner KS, Kondolf GM. 2002. <strong>River</strong>s and streams.<br />

In: Perrow MR, Davy AJ (Eds), Handbook of Ecological <strong>Restoration</strong>,<br />

Volume 2 <strong>Restoration</strong> <strong>in</strong> Practice. Cambridge University<br />

Press: Cambridge; 267–296.<br />

EEC. 1992. Directive 92/43/EEC of <strong>the</strong> European Parliament and<br />

of <strong>the</strong> Council of 21 May 1992 on <strong>the</strong> conservation of natural<br />

habitats and of wild fauna and fl ora. OJ L 206, 22 July 1992,<br />

The European Parliament: Brussels, Belgium.<br />

Environment Agency. 2002. A Framework for Catchment Abstraction<br />

Management Strategies. R&D Technical Report W6-066/<br />

TR, RAM version 3 Summary Report. Environment Agency:<br />

Bristol.<br />

European Union. 2000. Directive 2000/60/EC of <strong>the</strong> European<br />

Parliament and of <strong>the</strong> Council of 23 October 2000 Establish<strong>in</strong>g


102 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

a Framework for Community Action <strong>in</strong> <strong>the</strong> Field of Water<br />

Policy. OJ L 327, 23 October 2000, The European Parliament:<br />

Brussels, Belgium.<br />

Frissell CA, Nawa R. 1992. Incidences and causes of physical<br />

failure of artifi cial habit structures <strong>in</strong> streams of Western<br />

Oregon and Wash<strong>in</strong>gton. North American Journal of Fisheries<br />

Management 12: 187–197.<br />

Gibert J, Fournier F, Mathieu J. 1997. The groundwater/surface<br />

waters ecotone perspective: State of <strong>the</strong> art. In: Gibert J,<br />

Mathieu J, Fournier F (Eds), Groundwater/Surface Water Ecotones:<br />

Biological and Hydrological Interactions and Management<br />

Options. International Hydrology Series, Cambridge<br />

University Press: Cambridge; 3–8.<br />

Gillilan S, Boyd K, Hoitsma T, Kauffman M. 2005. Challenges<br />

<strong>in</strong> develop<strong>in</strong>g and implement<strong>in</strong>g ecological standards for geomorphic<br />

river restoration projects: A comment on Palmer et al.<br />

(2004). Journal of Applied Ecology 42 (2): 223–227.<br />

Girel J, Hughes FMR, Moss T et al. 2003. A case for fl oodpla<strong>in</strong><br />

forests? In: Hughes FMR. (Ed) The Flooded Forest: Guidance<br />

for Policy Makers and <strong>River</strong> Managers <strong>in</strong> Europe on <strong>the</strong> <strong>Restoration</strong><br />

of Floodpla<strong>in</strong> Forests. FLOBAR2, Department of<br />

Geography, University of Cambridge: Cambridge; 6–23.<br />

Godw<strong>in</strong> H. 1941. Studies of <strong>the</strong> post-glacial history of British<br />

Vegetation. II Fenland Pollen Diagrams. Philosophical Transactions<br />

of <strong>the</strong> Royal Society Series B 230: 239–284.<br />

Gopal B, Junk WJ. 2001. Assessment, determ<strong>in</strong>ants, function and<br />

conservation of biodiversity <strong>in</strong> wetlands: present status and<br />

future needs. In: Gopal B, Junk WJ, Davis JA (Eds), Biodiversity<br />

<strong>in</strong> Wetlands: Assessment, Conservation and Function,<br />

Volume 2. Backhuys Publishers: Leiden, The Ne<strong>the</strong>rlands;<br />

277–302.<br />

Gren I-M, Groth K-H, Sylvén M. 1995. Economic Values of<br />

Danube Floodpla<strong>in</strong>s. Journal of Environmental Management<br />

45: 333–345.<br />

Hancock PJ, Boulton AJ, Humphreys WF. 2005. Aquifers and<br />

hyporheic zones: Towards an ecological understand<strong>in</strong>g of<br />

groundwater. Hydrogeology Journal 13: 98–111.<br />

Haraszthy L. 2001. The fl oodpla<strong>in</strong> forests <strong>in</strong> Hungary. In: Klimo<br />

E, Hager H (Eds), The Floodpla<strong>in</strong> Forests <strong>in</strong> Europe: Current<br />

Situation and Perspectives. Kon<strong>in</strong>klijke Brill NV, Leiden and<br />

European Forest Institute Research Report No. 10: Leiden, The<br />

Ne<strong>the</strong>rlands; 17–24.<br />

Harris RR. 1999. Defi n<strong>in</strong>g reference conditions for restoration of<br />

riparian plant communities: examples from California, USA.<br />

Environmental Management 24: 55–63.<br />

Heywood VH (Ed). 1995. Global Biodiversity Assessment. Cambridge<br />

University Press: Cambridge.<br />

Hill M, Platts WS. 1991. Ecological and geomorphological concepts<br />

for <strong>in</strong>stream and out-of-channel fl ow requirements.<br />

<strong>River</strong>s 2: 319–343.<br />

Holl<strong>in</strong>g CS. 1973. Resilience and stability of ecological systems.<br />

Annual Review of Ecology and Systematics 5: 25–37.<br />

Hughes FMR. 1994. Environmental change, disturbance and<br />

regeneration <strong>in</strong> semi-arid fl oodpla<strong>in</strong> forests. In: Mill<strong>in</strong>gton AC,<br />

Pye K (Eds), Environmental Change <strong>in</strong> Drylands: Biogeographical<br />

and Geomorphological Perspectives. John Wiley &<br />

Sons Ltd: Chichester; 321–346.<br />

Hughes FMR, Muller E. 2003. How can fl oodpla<strong>in</strong> forests be<br />

restored. In: Hughes FMR (Ed), The Flooded Forest: Guidance<br />

for Policy Makers and <strong>River</strong> Managers <strong>in</strong> Europe on <strong>the</strong> <strong>Restoration</strong><br />

of Floodpla<strong>in</strong> Forests. FLOBAR2, Department of<br />

Geography, University of Cambridge: Cambridge; 49–70.<br />

Hughes FMR, Rood SB. 2001. Floodpla<strong>in</strong>s. In: Warren A, French<br />

JR (Eds), Habitat Conservation: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> Physical Environment<br />

John Wiley & Sons Ltd: Chichester; 105–121.<br />

Hughes FMR, Rood SB. 2003. The allocation of river fl ows for<br />

<strong>the</strong> restoration of woody riparian and fl oodpla<strong>in</strong> forest ecosystems:<br />

a review of approaches and <strong>the</strong>ir application <strong>in</strong> Europe.<br />

Environmental Management 32: 12–33.<br />

Hughes FMR, Colston A, Mountford JO. 2005. Restor<strong>in</strong>g riparian<br />

ecosystems: <strong>the</strong> challenge of accommodat<strong>in</strong>g variability and<br />

design<strong>in</strong>g restoration trajectories. Ecology and Society 10(1),<br />

article 12 Onl<strong>in</strong>e URL: http://www.ecologyandsociety.org/<br />

vol10/iss1/art12/<br />

Hulse DW, Gregory SV. 2004. Integrat<strong>in</strong>g resilience <strong>in</strong>to fl oodpla<strong>in</strong><br />

restoration. Urban Ecosystems 7: 295–314.<br />

Hulse DW, Gregory SV, Baker JP (Eds), 2002. Willamette <strong>River</strong><br />

Bas<strong>in</strong>: Trajectories of Environmental and Ecological Change.<br />

Oregon State University Press: Corvallis, Oregon.<br />

Jansson R, Backx H, Boulton AJ et al. 2005. Stat<strong>in</strong>g mechanisms<br />

and refi n<strong>in</strong>g criteria for ecologically successful river restoration:<br />

A comment on Palmer et al. (2005). Journal of Applied<br />

Ecology 42 (2): 218–222.<br />

Jensen MN. 2004. Climate warm<strong>in</strong>g shakes up species. BioScience<br />

54: 722–729.<br />

JNCC. 2004. Common Standard Monitor<strong>in</strong>g Guidance for <strong>River</strong>s.<br />

Jo<strong>in</strong>t Nature conservation Committee: Peterborough, UK.<br />

Johnson WC. 2000. Tree recruitment and survival <strong>in</strong> rivers: <strong>in</strong>fl uence<br />

of hydrological processes. Hydrological Processes 14:<br />

3051–3074.<br />

Junk WJ, Bayley PB, Sparks RE. 1989. The fl ood-pulse concept<br />

<strong>in</strong> river-fl oodpla<strong>in</strong> systems. Canadian Special Publication of<br />

Fisheries and Aquatic Sciences 106: 110–127.<br />

K<strong>in</strong>g J, Louwe D. 1998. Instream fl ow requirements for regulated<br />

rivers <strong>in</strong> South Africa us<strong>in</strong>g <strong>the</strong> build<strong>in</strong>g block methodology.<br />

Aquatic Ecosystem Health and Management 1: 109–124.<br />

K<strong>in</strong>g J, Brown C, Sabat H. 2003. A scenario-based holistic<br />

approach to environmental fl ow assessments for rivers. <strong>River</strong><br />

Research and Applications 19: 619–639.<br />

Lambs L. 2004. Interactions between groundwater and surface<br />

water at river banks and <strong>the</strong> confl uence of rivers. Journal of<br />

Hydrology 288: 312–326.<br />

MacNally R, Park<strong>in</strong>son A, Horrocks G, Young M. 2002. Current<br />

loads of coarse woody debris on sou<strong>the</strong>astern Australian fl oodpla<strong>in</strong>s:<br />

Evaluation and change and implications for restoration.<br />

Ecological <strong>Restoration</strong> 10 (4): 627–635.<br />

Mahoney JM, Rood SB. 1998. Streamfl ow requirements for cottonwood<br />

seedl<strong>in</strong>g recruitment-An <strong>in</strong>tegrative Model. Wetlands<br />

18: 634–645.<br />

Malanson CP. 1993. Riparian Landscapes Cambridge University<br />

Press: Cambridge.<br />

Marston RA, Girel J, Pautou G et al. 1995. Channel metamorphosis,<br />

fl oodpla<strong>in</strong> disturbance, and vegetation development –<br />

A<strong>in</strong> <strong>River</strong>. Geomorphology 13: 121–131.


May RM. 1976a. Models for two <strong>in</strong>teract<strong>in</strong>g populations. In: May<br />

RM (Ed.), Theoretical Ecology: Pr<strong>in</strong>ciples and Applications.<br />

Blackwell Scientifi c Publications: Oxford; 49–70.<br />

May RM. 1976b. Patterns <strong>in</strong> multi-species communities. In: May<br />

RM (Ed.), Theoretical Ecology: Pr<strong>in</strong>ciples and Applications.<br />

Blackwell Scientifi c Publications: Oxford; 142–162.<br />

Medley K, Hughes FMR. 1996. <strong>River</strong><strong>in</strong>e forests. In: McClanahan<br />

TR (Ed.), Ecosystems and <strong>the</strong>ir Conservation <strong>in</strong> East Africa.<br />

Oxford University Press: Oxford; 361–384.<br />

Michelot J-L. 1995. Gestion Patrimoniale des Milieux Naturels<br />

Fluviaux: Guide Technique. Reserves Naturelles de France.<br />

Middelkoop H, van Haselen COG (Eds), 1999. Twice a <strong>River</strong> –<br />

Rh<strong>in</strong>e and Meuse <strong>in</strong> <strong>the</strong> Ne<strong>the</strong>rlands. RIZA Report no. 99.003,<br />

Institute for Inland Water Management and Waste Water Treatment/RIZA:<br />

Arnhem, The Ne<strong>the</strong>rlands.<br />

Middleton B. 1999. Flood Puls<strong>in</strong>g and Disturbance Dynamics.<br />

John Wiley & Sons Inc.: New York.<br />

Montgomery DR. 1999. Process doma<strong>in</strong>s and <strong>the</strong> river cont<strong>in</strong>uum.<br />

Journal of <strong>the</strong> American Water Resources Association 35:<br />

397–410.<br />

Montgomery DR, Boulton SM. 2003. In: Wissmar RC, Bisson<br />

PA (Eds), Strategies for Restor<strong>in</strong>g <strong>River</strong> Ecosystems: Sources<br />

of Variability and <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Natural and Managed Systems.<br />

American Fisheries Society: Be<strong>the</strong>sda, Maryland; 39–80.<br />

Moss T. 2003. Solv<strong>in</strong>g problems of ‘fi t’ at <strong>the</strong> expense of problems<br />

of ‘<strong>in</strong>terplay’? The spatial reorganisation of water management<br />

follow<strong>in</strong>g <strong>the</strong> EU Water Framework Directive. In:<br />

Breit H, Engels A, Moss T, Troja M (Eds), How Institutions<br />

Change. Perspectives on Social Learn<strong>in</strong>g <strong>in</strong> Global and Local<br />

Environmental Contexts. Leske and Budrich: Opladen,<br />

Germany; 85–121.<br />

Naiman RJ, Bunn SE, Nilsson C et al. 2002. Legitimiz<strong>in</strong>g fl uvial<br />

ecosystems as users of water: An overview. Environmental<br />

Management 30: 455–467.<br />

Nilsson C, Grelsson G, Johansson M, Sperens U. 1991a. Small<br />

rivers behave like large rivers: effects of postglacial history on<br />

plant species richness along river banks. Journal of Biogeography<br />

18: 533–541.<br />

Nilsson C, Ekblad A, Gardfjell M, Carlberg B. 1991b. Long term<br />

effects of river regulation on river marg<strong>in</strong> vegetation. Journal<br />

of Applied Ecology 28: 963–987.<br />

Nilsson C, Lepori F, Malmqvist B et al. 2005. Forecast<strong>in</strong>g<br />

environmental responses to restoration of rivers used as log<br />

fl oatways: An <strong>in</strong>terdiscipl<strong>in</strong>ary challenge. Ecosystems 8:<br />

779–800.<br />

NRC (National Research Council). 1992. New Strategies for<br />

America’s Watersheds. National Academic Press: Wash<strong>in</strong>gton,<br />

DC.<br />

Olden JD, Poff NL. 2003. Redundancy and <strong>the</strong> choice of hydrologic<br />

<strong>in</strong>dices for characteriz<strong>in</strong>g streamfl ow regimes. <strong>River</strong><br />

Research and Applications 19: 101–121.<br />

Pahl-Wostl C. 2002. Participative and stakeholder-based policy<br />

design, analysis and evaluation processes. Integrated Assessment<br />

3: 3–14.<br />

Pahl-Wostl C. 2004. Information, public empowerment, and <strong>the</strong><br />

management of urban watersheds. Environmental Modell<strong>in</strong>g<br />

and Software 20: 457–467.<br />

<strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Riparian and Floodpla<strong>in</strong> <strong>Restoration</strong> 103<br />

Palmer MA, Bernhardt ES, Allen JD et al. 2005. Standards for<br />

ecologically successful river restoration. Journal of Applied<br />

Ecology 42 (2): 208–217.<br />

Petts GE. 1990. The role of ecotones <strong>in</strong> aquatic landscape management.<br />

In: Naiman RJ, Decamps H (Eds), The Ecology and<br />

Management of Aquatic-Terrestrial Ecotones. UNESCO/MAB<br />

Series 4: Par<strong>the</strong>non, New Jersey; 227–261.<br />

Petts GE, Gurnell AM, Gerrard AJ et al. 2000. Longitud<strong>in</strong>al<br />

variations <strong>in</strong> exposed river<strong>in</strong>e sediments; a context for <strong>the</strong><br />

ecology of <strong>the</strong> Fiume Tagliamento, Italy. Aquatic Conservation:<br />

Mar<strong>in</strong>e and Freshwater Ecosystems 10: 249–266.<br />

Planty-Tabacchi A-M, Tabacchi E, Naiman RJ et al. 1996. Invasibility<br />

of species-rich communities <strong>in</strong> riparian zones. Conservation<br />

Biology 10: 598–607.<br />

Poff NL. 1992. Why disturbance can be predictable: A perspective<br />

on <strong>the</strong> defi nition of disturbance <strong>in</strong> streams. Journal of <strong>the</strong> North<br />

American Benthological Society 11: 86–92<br />

Poff NL, Allan JD, Palmer MA et al. 2003. <strong>River</strong> fl ows and water<br />

wars: emerg<strong>in</strong>g science for environmental decision mak<strong>in</strong>g.<br />

Frontiers <strong>in</strong> Ecology and Environment 1 (6): 298–306.<br />

Postel S, Richter B. 2003. <strong>River</strong>s for Life: <strong>Manag<strong>in</strong>g</strong> Water for<br />

People and Nature. Island Press: Wash<strong>in</strong>gton, DC.<br />

Power M. 1997. The Audit Society. Rituals of Verifi cation. Oxford<br />

University Press: Oxford.<br />

Richards KS, Bras<strong>in</strong>gton J, Hughes FMR. 2002. Geomorphic<br />

dynamics of fl oodpla<strong>in</strong>s: Ecological implications and a potential<br />

modell<strong>in</strong>g strategy. Freshwater Biology 47: 1–22.<br />

Richter B, Postel S. 2004. Sav<strong>in</strong>g Earth’s rivers. Issues <strong>in</strong> Science<br />

and Technology Spr<strong>in</strong>g issue: 31–36.<br />

Richter BD, Ma<strong>the</strong>ws R, Harrison DL, Wig<strong>in</strong>gton R. 2003. Ecologically<br />

susta<strong>in</strong>able water management: manag<strong>in</strong>g river fl ows<br />

for ecological <strong>in</strong>tegrity. Ecological Applications 13 (1):<br />

206–224.<br />

Rodwell JS (Ed.). 1991a. British Plant Communities, Volume 1:<br />

Woodlands and Scrub. Cambridge University Press:<br />

Cambridge.<br />

Rodwell JS (Ed.). 2000. British Plant Communities, Volume 2:<br />

Mires and Heath. Cambridge University Press: Cambridge.<br />

Rodwell JS (Ed.). 1991b. British Plant Communities, Volume 3:<br />

Grasslands and Montane Communities. Cambridge University<br />

Press: Cambridge.<br />

Rodwell JS (Ed.). 1992. British Plant Communities, Volume 4:<br />

Aquatic Communities, Swamps and Tall-Herb Fens. Cambridge<br />

University Press: Cambridge.<br />

Rodwell JS (Ed.). 1995. British Plant Communities, Volume 5:<br />

Maritime Communities and Vegetation of Open Habitats. Cambridge<br />

University Press: Cambridge.<br />

Rood SB, Mahoney JM. 2000. Revised <strong>in</strong>stream fl ow regulation<br />

enables cottonwood recruitment along <strong>the</strong> St Mary <strong>River</strong>,<br />

Alberta, Canada. <strong>River</strong>s 7: 109–125.<br />

Rood SB, Mahoney JM, Reid D, Zilm L. 1995. Instream<br />

fl ows and <strong>the</strong> decl<strong>in</strong>e of riparian cottonwoods along <strong>the</strong><br />

St. Mary <strong>River</strong>, Alberta. Canadian Journal of Botany 73:<br />

1250–1260.<br />

Rood SB, Gourley C, Ammon EM et al. 2003. Flows for fl oodpla<strong>in</strong><br />

forests: successful riparian restoration along <strong>the</strong> lower<br />

Truckee <strong>River</strong>, Nevada, U.S.A. BioScience 53: 647–656.


104 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

RSPB, NRA and RSNC. 1994. The New <strong>River</strong>s and Wildlife<br />

Handbook. RSPB: Sandy, Bedfordshire.<br />

Salo J, Kalliola R, Häkk<strong>in</strong>en I et al. 1986. <strong>River</strong> dynamics and<br />

<strong>the</strong> diversity of Amazon lowland forests. Nature 322:<br />

254–258.<br />

Scheimer F, Baumgartner C, Tockner K. 1999. <strong>Restoration</strong> of<br />

fl oodpla<strong>in</strong> rivers: <strong>the</strong> Danube restoration project. Regulated<br />

<strong>River</strong>s, Research and Management 15: 231–244.<br />

Simons J, Boeters R. 1998. A systematic approach to ecologically<br />

sound river bank management. In: de Waal LC, Large ARG,<br />

Wade PM (Eds), Rehabilitation of <strong>River</strong>s: Pr<strong>in</strong>ciples and<br />

Implementation. John Wiley & Sons Ltd: Chichester; 57–85.<br />

Stream Corridor Work<strong>in</strong>g Group. 1998. Stream Corridor <strong>Restoration</strong>:<br />

Pr<strong>in</strong>ciples, Processes and Practices. Federal Stream<br />

Interagency Stream <strong>Restoration</strong> Work<strong>in</strong>g Group, USA.<br />

Tabacchi E, Planty-Tabacchi AM, Sal<strong>in</strong>as MJ, Decamps H. 1996.<br />

Landscape structure and diversity <strong>in</strong> riparian plant communities:<br />

A longitud<strong>in</strong>al comparative study. Regulated <strong>River</strong>s:<br />

Research and Management 12: 367–390.<br />

Tockner K, Malard F, Ward JV. 2000. An extension of <strong>the</strong> Flood<br />

Pulse concept. Hydrological Processes 14: 2861–2883.<br />

Turner RK, van den Bergh JCJM, Söderqvist T et al. 2000. Ecological-economic<br />

analysis of wetlands: scientifi c <strong>in</strong>tegration<br />

for management and policy. Ecological Economics 35: 7–23.<br />

UNEP – World Conservation Monitor<strong>in</strong>g Centre. 2000. European<br />

Forests and Protected Areas: Gap Analysis. UNEP – WCMC:<br />

Cambridge, UK.<br />

Vannote RL, M<strong>in</strong>shall GW, Cumm<strong>in</strong>s KW et al. 1980. The river<br />

cont<strong>in</strong>uum concept. Canadian Journal of Fisheries and Aquatic<br />

Sciences 37: 130–137.<br />

Ward JV, Stanford JA. 1995. Ecological connectivity <strong>in</strong><br />

alluvial river ecosystems and its disruption by fl ow regulation.<br />

Regulated <strong>River</strong>s: Research and Management 11: 105–<br />

119.<br />

Ward JV, Tockner K. 2001. Biodiversity: towards a unify<strong>in</strong>g<br />

<strong>the</strong>me for river ecology. Freshwater Biology 46: 807–819.<br />

Ward JV, Tockner K, Schmeier F. 1999. Biodiversity of fl oodpla<strong>in</strong><br />

ecosystems: ecotones and connectivity. Regulated <strong>River</strong>s:<br />

Research & Management 15: 125–139.<br />

Ward JV, Tockner K, Arscott DB, Claret C. 2002. <strong>River</strong><strong>in</strong>e landscape<br />

diversity. Freshwater Biology 47: 517–539.<br />

War<strong>in</strong>g RH. 1989. Ecosystems: fl uxes of matter and energy. In:<br />

Cherrett JM (Ed.), Ecological Concepts: The Contribution of<br />

Ecology to an Understand<strong>in</strong>g of <strong>the</strong> Natural World. British<br />

Ecological Society and Blackwell Scientifi c Publications:<br />

Oxford; 17–42.<br />

World Commission on Dams 2000. Dams and Development: A<br />

New Framework for Decision Mak<strong>in</strong>g. Earthscan Publications<br />

Ltd: London.<br />

WWF 2000. Wise Use of Floodpla<strong>in</strong>s. Policy and Economic<br />

Analysis of Floodpla<strong>in</strong> <strong>Restoration</strong> <strong>in</strong> Europe. Opportunities<br />

and Obstacles. Report, November 2000.<br />

WWF 2004. Liv<strong>in</strong>g with Floods: Achiev<strong>in</strong>g Ecologically Susta<strong>in</strong>able<br />

Flood Management <strong>in</strong> Europe. Report, June 2004.<br />

Young O. 1999. Institutional Dimensions of Global Environmental<br />

Change. Science Plan. IHDP Report No. 9, IHDP: Bonn,<br />

Germany.<br />

Young O. 2002. The Institutional Dimensions of Environmental<br />

Change. Fit, Interplay, and Scale. MIT Press: Cambridge,<br />

Massachusetts.<br />

Zöckler C. 2000. Wise Use of Floodpla<strong>in</strong>s – LIFE Environment<br />

Project: A Review of 12 WWF <strong>River</strong> <strong>Restoration</strong> Projects<br />

Across Europe. WWF European Freshwater Programme:<br />

Copenhagen, Denmark.


<strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat Edited by Stephen Darby and David Sear<br />

© 2008 John Wiley & Sons, Ltd<br />

7<br />

Hydrological and Hydraulic Aspects of<br />

<strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for<br />

Ecological Purposes<br />

7.1 INTRODUCTION<br />

N.J. Clifford 1 , M.C. Acreman 2 and D.J. Booker 2,3<br />

1 School of Geography, University of Nott<strong>in</strong>gham, UK<br />

2 Centre for Ecology and Hydrology, UK<br />

3 (now at) National Institute of Water and Atmospheric Research, New Zealand<br />

This chapter presents river restoration as a hybrid activity,<br />

<strong>in</strong>volv<strong>in</strong>g hydrological, geomorphological and ecological<br />

expertise. It <strong>in</strong>troduces a range of techniques and methods<br />

that are appropriate to each of <strong>the</strong>se areas, gives examples<br />

of <strong>the</strong>ir application and reviews some of <strong>the</strong> fundamental<br />

opportunities and limitations of current restoration practice.<br />

<strong>River</strong> restoration is not only a hybrid activity but also<br />

an emerg<strong>in</strong>g one (both scientifi cally and practically).<br />

‘Uncerta<strong>in</strong>ties’ are present at every stage of restoration<br />

<strong>in</strong>tervention. These span such basic issues as: <strong>the</strong> ability<br />

to support ideas of catchment hydrology or fl ow regime<br />

with<strong>in</strong> which to frame restoration designs; provid<strong>in</strong>g fi eld<br />

evidence for conceptual and numerical simulation models;<br />

captur<strong>in</strong>g <strong>the</strong> natural range of variability <strong>in</strong>herent <strong>in</strong><br />

complex and dynamic physical–ecological systems; and<br />

<strong>in</strong>corporat<strong>in</strong>g all of <strong>the</strong>se uncerta<strong>in</strong>ties <strong>the</strong>mselves <strong>in</strong>to<br />

restoration design and appraisal practice. The chapter concludes<br />

that fi ve basic sources of uncerta<strong>in</strong>ty (also see<br />

Chapters 1 to 3) underp<strong>in</strong> contemporary river restoration<br />

<strong>the</strong>ory and practice: data (type, quality and quantity);<br />

characterisation (of those physical and ecological phenomena<br />

<strong>in</strong>volved <strong>in</strong> <strong>the</strong> restoration attempt); coupl<strong>in</strong>g (of<br />

physical environmental and ecological dynamics); awareness<br />

(of opportunities and limitations) and fl exibility (<strong>in</strong><br />

<strong>the</strong> approach to design and evaluation). Evaluat<strong>in</strong>g each<br />

of <strong>the</strong>se sources at every stage of <strong>the</strong> restoration is,<br />

perhaps, <strong>the</strong> best way of manag<strong>in</strong>g such uncerta<strong>in</strong>ties and,<br />

too, of improv<strong>in</strong>g prospects for <strong>the</strong> <strong>in</strong>corporation of more<br />

sophisticated simulation approaches to river restoration<br />

design and appraisal <strong>in</strong> <strong>the</strong> future.<br />

7.2 THE RIVER AND CATCHMENT AS AN<br />

UNCERTAIN SYSTEM<br />

<strong>Restoration</strong> of rivers may be viewed as <strong>the</strong> jo<strong>in</strong>t product<br />

of <strong>the</strong> physical structure of <strong>the</strong> channel and its fl oodpla<strong>in</strong>,<br />

<strong>the</strong>ir hydrological <strong>in</strong>tegration and <strong>the</strong>ir ecological value<br />

and function. Channels and fl oodpla<strong>in</strong>s are characterised<br />

by <strong>the</strong> tim<strong>in</strong>g and quantity of fl ows and sediments which<br />

<strong>the</strong> channel conveys, and which <strong>the</strong> fl oodpla<strong>in</strong> stores. As<br />

a result, channel restoration <strong>in</strong> its widest sense encompasses<br />

structural modifi cations of channel form, <strong>the</strong> reestablishment<br />

of natural fl ow regime and <strong>the</strong> reconnection<br />

(or even recreation) of channel and fl oodpla<strong>in</strong> areas (see<br />

Chapter 6). Both <strong>the</strong> problems of, and solution to, hydrological<br />

and hydraulic uncerta<strong>in</strong>ties <strong>in</strong> river restoration thus<br />

arise from consider<strong>in</strong>g <strong>the</strong> channel as an embedded part<br />

of <strong>the</strong> wider fl uvial hydrosystem (Petts and Amoros,<br />

1996). Incorporat<strong>in</strong>g hydrological connections or disturbances<br />

requires an holistic approach to streamfl ow management<br />

(Hill et al., 1991) <strong>in</strong> recognition of <strong>the</strong> many<br />

scales of connection or ‘dimensions’ of <strong>the</strong> fl uvial system<br />

(Boon, 1998; and for general review see Clifford 2001).<br />

Approaches to cost effective and multifunction river<br />

rehabilitation works have <strong>in</strong>creas<strong>in</strong>gly emphasised <strong>the</strong><br />

need to <strong>in</strong>clude an ecological perspective. Attention<br />

has focused on restor<strong>in</strong>g susta<strong>in</strong>able hydrological and


106 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

hydraulic habitats us<strong>in</strong>g pr<strong>in</strong>ciples of fl uvial behaviour<br />

(Newbury and Gaboury, 1993) as shown <strong>in</strong> Figure 7.1.<br />

This fi gure illustrates <strong>the</strong> hierarchy of feedbacks (or <strong>in</strong>terconnections)<br />

between physical structure, hydrological<br />

function and ecological responses <strong>in</strong> an alluvial river–<br />

fl oodpla<strong>in</strong> system which are relevant when design<strong>in</strong>g and<br />

assess<strong>in</strong>g river restoration or rehabilitation schemes. Flow<br />

and sediment transport depend upon catchment <strong>in</strong>puts of<br />

material and energy. These <strong>in</strong> turn ma<strong>in</strong>ta<strong>in</strong> and confi gure<br />

<strong>the</strong> channel shape, determ<strong>in</strong>e <strong>the</strong> sediment and bedform<br />

environments and create diversity of fl ow patterns and<br />

structures with vary<strong>in</strong>g degrees of coherence and spatial<br />

coverage. Ecological function of <strong>the</strong> channel is primarily<br />

a response to local (imposed) conditions of velocity, depth<br />

and substrate (<strong>the</strong> hydraulic variables), whose spatial<br />

(cross-section, reach-scale) and temporal (event-specifi c<br />

Human Land<br />

Use and Flow<br />

Regulation<br />

• water<br />

• sediment<br />

• nutrients<br />

Watershed Inputs<br />

Fluvial Geomorphic Processes<br />

• sediment transport/deposition/scour<br />

• channel migration and bank erosion<br />

• floodpla<strong>in</strong> construction and <strong>in</strong>undation<br />

• surface and groundwater <strong>in</strong>teractions<br />

Geomorphic Attributes<br />

• channel morphology (size, slope, shape,<br />

bed and bank composition)<br />

• floodpla<strong>in</strong> morphology<br />

• water turbidity and temperature<br />

Habitat Structure, Complexity, and Connectivity<br />

• <strong>in</strong>stream aquatic habitat<br />

• shaded riparian aquatic habitat<br />

• riparian woodlands<br />

• seasonally <strong>in</strong>undated floodpla<strong>in</strong> wetlands<br />

Biotic Responses<br />

(Aquatic, Riparian, and Terrestrial Plants and Animals)<br />

• abundance and distribution of native and exotic species<br />

• community composition and structure<br />

• food web structure<br />

and seasonal) characteristics refl ect wider reach-scale,<br />

<strong>in</strong>ter-reach and catchment controls (which are primarily<br />

hydrologically determ<strong>in</strong>ed).<br />

While <strong>the</strong> nature of <strong>the</strong> l<strong>in</strong>kages <strong>in</strong> Figure 7.1 is well<br />

understood, giv<strong>in</strong>g precise values to quantities and tim<strong>in</strong>gs<br />

of material and energy transfers, and account<strong>in</strong>g for <strong>the</strong><br />

feedbacks between <strong>the</strong>m, gives rise to uncerta<strong>in</strong>ties at all<br />

scales. These uncerta<strong>in</strong>ties are compounded by <strong>the</strong> recognition<br />

that climate and land use, which determ<strong>in</strong>e catchment<br />

ra<strong>in</strong>fall–run-off response, are nonstationary (that is,<br />

<strong>the</strong>y are chang<strong>in</strong>g <strong>in</strong> <strong>the</strong>ir mean level and variance) through<br />

time. In <strong>the</strong> United K<strong>in</strong>gdom, for example, Prudhomme<br />

et al. (2003) exam<strong>in</strong>e <strong>the</strong> implications of no less than<br />

25 000 climate scenarios for four typical fl ood events as<br />

applied to fi ve catchments. Most scenarios show an<br />

<strong>in</strong>crease <strong>in</strong> both <strong>the</strong> magnitude and frequency of fl ood<br />

• energy<br />

• large woody debris<br />

• chemical pollutants<br />

Natural<br />

Disturbance<br />

Figure 7.1 The nested hierarchy of <strong>the</strong> channel system and its associated habitat potential and biotic response (Tuolumne <strong>River</strong><br />

restoration program summary report, summary of studies, conceptual models, restoration projects, and ongo<strong>in</strong>g monitor<strong>in</strong>g. Prepared<br />

for <strong>the</strong> CALFED/AFRP Adaptive Management Forum, with assistance from <strong>the</strong> Tuolun<strong>in</strong>e <strong>River</strong> Technical Advisory Committee,<br />

(2001). Reproduced with permission from Stillwater Sciences.)


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 107<br />

events, but <strong>the</strong> largest uncerta<strong>in</strong>ty arises from <strong>the</strong> type<br />

of Global Climate Model used: <strong>the</strong> magnitude of modelled<br />

change varies by a factor of n<strong>in</strong>e <strong>in</strong> Nor<strong>the</strong>rn England<br />

and Scotland! At scales below <strong>the</strong> level of climatic<br />

<strong>in</strong>put, major uncerta<strong>in</strong>ties rema<strong>in</strong> <strong>in</strong> <strong>the</strong> modell<strong>in</strong>g and<br />

prediction of ra<strong>in</strong>fall–run-off relationships, upon which<br />

channel fl ow ultimately depends. Catchment run-off modell<strong>in</strong>g<br />

is <strong>in</strong>creas<strong>in</strong>gly able to <strong>in</strong>corporate greater physical<br />

parameterisation, and to distribute parameter comb<strong>in</strong>ations<br />

around <strong>the</strong> catchment and between surface and subsurface<br />

fl ows. Similar predictions may, however, be<br />

obta<strong>in</strong>ed us<strong>in</strong>g models that <strong>in</strong>corporate differ<strong>in</strong>g levels of<br />

sophistication and differ<strong>in</strong>g parameter values, giv<strong>in</strong>g rise<br />

to model ‘equifi nality’. This raises concerns about <strong>the</strong><br />

degree to which fundamental understand<strong>in</strong>g of process<br />

can, <strong>in</strong> fact, lead to better <strong>in</strong>formation on future (or<br />

designed) outcomes (Beven, 2001). Ano<strong>the</strong>r major source<br />

of concern is <strong>the</strong> degree to which model predictions may<br />

be appropriately up- or down-scaled (for a comprehensive<br />

review of <strong>the</strong>se issues, see Beven, 2000 and Bierkens<br />

et al., 2000).<br />

Gilvear et al. (2002) and Hendry et al. (2003) po<strong>in</strong>t out<br />

that <strong>the</strong> hydrological basis for determ<strong>in</strong><strong>in</strong>g future fi sheries<br />

stocks is complicated by issues of water and sediment<br />

quality, as well as quantity. Water quality necessitates<br />

consideration of longer- and shorter-term land use histories,<br />

run-off behaviour and <strong>the</strong> monitor<strong>in</strong>g or modell<strong>in</strong>g<br />

of diffuse as well as po<strong>in</strong>t-source pollutants. ‘Hydrology’<br />

itself, <strong>the</strong>refore, has many uncerta<strong>in</strong> components,<br />

and Clarke SL et al. (2003) call for <strong>the</strong> development<br />

of ‘hydrogeomorphological knowledge’ of catchments<br />

support<strong>in</strong>g ‘tools’ for water resource management. The<br />

need to service ever-more complex models and to comply<br />

with <strong>in</strong>creas<strong>in</strong>gly complex (but frequently compet<strong>in</strong>g) legislative<br />

requirements will also place grow<strong>in</strong>g pressures for<br />

<strong>the</strong> provision of data (both quantity and quality). This is<br />

likely to demand changes to long-stand<strong>in</strong>g methods of<br />

data acquisition and process<strong>in</strong>g (Marsh, 2002). All of<br />

<strong>the</strong>se essentially hydrological issues are encountered<br />

before <strong>the</strong> channel-scale is addressed from a hydraulic<br />

standpo<strong>in</strong>t!<br />

With<strong>in</strong> <strong>the</strong> channel at reach and subreach scales, it is<br />

hydraulics which underp<strong>in</strong> connections between <strong>the</strong> physical<br />

and ecological environments. Determ<strong>in</strong><strong>in</strong>g an appropriate<br />

channel morphology and <strong>the</strong> characteristics of fl ow<br />

behaviour <strong>in</strong> response to chang<strong>in</strong>g discharge and sediment<br />

transport are key factors <strong>in</strong> design<strong>in</strong>g restoration schemes.<br />

Such schemes must be both susta<strong>in</strong>able <strong>in</strong> a physical<br />

sense, and functional <strong>in</strong> an ecological sense. Both considerations<br />

require some allowance for natural dynamics and<br />

post-design evolution. Over <strong>the</strong> last two decades, a new<br />

subject of ‘eco-hydraulics’ has developed (Leclerc, 2002),<br />

<strong>in</strong> which a range of monitor<strong>in</strong>g and modell<strong>in</strong>g strategies<br />

l<strong>in</strong>k <strong>the</strong> expertise of eng<strong>in</strong>eers, geomorphologies and<br />

ecologists to design and assess restoration or rehabilitation<br />

schemes. Yet, from very basic stages of fl ow ‘characterisation’<br />

through to fl ow modell<strong>in</strong>g, habitat simulation and<br />

post-project appraisal, numerous uncerta<strong>in</strong>ties exist. This<br />

chapter reviews <strong>the</strong> sources and implications of <strong>the</strong>se<br />

uncerta<strong>in</strong>ties, and provides case study examples of current<br />

and prospective restoration practice motivated by ecohydraulic<br />

considerations.<br />

7.3 HYDROLOGICAL ASPECTS AND<br />

UNCERTAINTY IN CHANNEL RESTORATION<br />

7.3.1 Connectivity, Disturbance and <strong>the</strong> Ecological<br />

Flow Regime <strong>in</strong> Channels<br />

A river ecosystem and its associated benefi ts to humank<strong>in</strong>d<br />

are strongly conditioned by <strong>the</strong> pattern of fl ows<br />

between days, seasons and years (<strong>in</strong>clud<strong>in</strong>g fl oods and<br />

droughts) that occur with<strong>in</strong> <strong>the</strong> dra<strong>in</strong>age bas<strong>in</strong>. For<br />

example, fl oods ma<strong>in</strong>ta<strong>in</strong> river structure and sediment<br />

distribution (Hill et al., 1991), medium fl ows trigger<br />

fi sh migration (Junk et al., 1989) and low fl ows ma<strong>in</strong>ta<strong>in</strong><br />

species diversity (Everard, 1996). The pattern of fl ows<br />

required to support a river ecosystem is called <strong>the</strong> environmental<br />

fl ow requirements (Dyson et al., 2003). If <strong>the</strong><br />

river fl ow pattern is altered from its natural regime, <strong>the</strong>n<br />

<strong>the</strong> river ecosystem will change from its natural state. Too<br />

much fl ow at <strong>the</strong> wrong time can be as damag<strong>in</strong>g as too<br />

little fl ow. The fl ow regime of a river describes <strong>the</strong> temporal<br />

variability of run-off with<strong>in</strong> a s<strong>in</strong>gle hydrological<br />

year (<strong>in</strong>ter-annual, such as between w<strong>in</strong>ter fl oods, spr<strong>in</strong>g<br />

snowmelt run-off, summer basefl ow), and from year-toyear<br />

(<strong>in</strong>tra-annual, such as dry years, wet years or alternat<strong>in</strong>g<br />

periods or cycles of drought and higher fl ows).<br />

Recognition of <strong>the</strong> importance of dynamism and adjustment<br />

<strong>in</strong> fl ow regime for <strong>the</strong> physical and biotic system is<br />

someth<strong>in</strong>g of a recent paradigm shift <strong>in</strong> river management<br />

and restoration (Bergen et al., 2001; Newson, 2002), but<br />

poses areas of additional design and management uncerta<strong>in</strong>ty.<br />

Traditional eng<strong>in</strong>eer<strong>in</strong>g <strong>in</strong>tervention was guided by<br />

pr<strong>in</strong>ciples of control and stability, which led to ‘certa<strong>in</strong>’<br />

or fi xed design and performance criteria. Design<strong>in</strong>g-<strong>in</strong><br />

variability, or allow<strong>in</strong>g for a degree of post-modifi cation<br />

‘naturalisation’ to channel works is not only a novel<br />

concept but is also largely untried or tested. It requires <strong>the</strong><br />

sett<strong>in</strong>g of generous tolerances on design functional requirements,<br />

to refl ect <strong>the</strong> degree of ignorance and paucity of<br />

research (Bergen et al., 2001). In this respect, ‘uncerta<strong>in</strong>ty’<br />

is itself someth<strong>in</strong>g of an essential design criterion<br />

(also see Chapter 14)!<br />

The most extreme changes to a fl ow regime are often<br />

associated with construction of a dam, where <strong>the</strong> entire


108 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

pattern of fl ows may be replaced by a constant low fl ow.<br />

This can have serious negative impacts on <strong>the</strong> river ecosystem<br />

downstream (Petts, 1984). A key part of river restoration<br />

is often <strong>the</strong> reduction of abstractions (Barker and<br />

Kirmond, 1998) or release of water from dams (Acreman,<br />

2002) to restore river fl ows to a more natural pattern.<br />

While geomorphologists and hydrologists have tended to<br />

emphasise connectivity and transfer of water and materials<br />

through <strong>the</strong> fl uvial system, <strong>the</strong> ‘real’ state of catchments<br />

<strong>in</strong> <strong>the</strong> developed world is of highly fragmented, and<br />

largely modifi ed transfers. Graf (2001) estimates that only<br />

2% of <strong>the</strong> 3.5 million miles of streams <strong>in</strong> <strong>the</strong> United States<br />

are unaffected by dams, and that 18% of this is actually<br />

under <strong>the</strong> waters of reservoirs. The picture of fragmentation,<br />

threat and change is similar <strong>in</strong> <strong>the</strong> United K<strong>in</strong>gdom,<br />

where less than 10% of rivers are free from structural<br />

modifi cation of channel and banks and 53% of rivers have<br />

fl ow regimes altered by more than 20% (Acreman, 2000).<br />

Indeed, such is <strong>the</strong> scale of <strong>in</strong>terference with natural conditions<br />

of catchments <strong>in</strong> developed countries, that <strong>the</strong> most<br />

conv<strong>in</strong>c<strong>in</strong>g strategy for environmental enhancement <strong>in</strong> <strong>the</strong><br />

short to medium term may be to use <strong>the</strong> capacity for<br />

regulation to remediate fl ows. The alternative, to remove<br />

expensive, fi xed <strong>in</strong>frastructure at <strong>the</strong> larger scale (i.e.<br />

restoration of fl ows), may require unrealistic cultural,<br />

political and economic shifts (Graf, 2001).<br />

Remediation of fl ows might be accomplished by changes<br />

to <strong>the</strong> operat<strong>in</strong>g rules of dams, by redesign and by physical<br />

renovation of structures. A high profi le case of relax<strong>in</strong>g<br />

impoundment occurred on <strong>the</strong> Colorado <strong>River</strong> <strong>in</strong> <strong>the</strong> United<br />

States <strong>in</strong> 1998 (Schmidt et al., 1998) and illustrates<br />

Pr<strong>in</strong>ciple 1<br />

variation <strong>in</strong> flow regime<br />

Habitat complexity biotic diversity<br />

discharge<br />

Pr<strong>in</strong>ciple 4<br />

time<br />

natural regime discourages <strong>in</strong>vasions<br />

<strong>the</strong> importance of unit<strong>in</strong>g o<strong>the</strong>rwise confl ict<strong>in</strong>g policies<br />

to achieve environmental benefi ts. Acreman (2002)<br />

addresses <strong>the</strong> implications of modifi ed natural systems for<br />

<strong>the</strong> practice of scientifi c hydrology, argu<strong>in</strong>g that <strong>the</strong> realities<br />

of modifi ed systems must be <strong>in</strong>corporated <strong>in</strong>to traditional<br />

hydrological tra<strong>in</strong><strong>in</strong>g and practice (for review see Clifford,<br />

2002).<br />

The natural fl ow regime of a river is a function of <strong>the</strong><br />

magnitude, duration, frequency and tim<strong>in</strong>g of precipitation;<br />

<strong>the</strong> form of <strong>the</strong> precipitation (ra<strong>in</strong> or snow) and <strong>the</strong><br />

characteristics of <strong>the</strong> dra<strong>in</strong>age bas<strong>in</strong> (which determ<strong>in</strong>es<br />

how precipitation translates <strong>in</strong>to streamfl ow via surface<br />

and subsurface run-off). Each stream has a unique fl ow<br />

regime, characterised by <strong>the</strong> stream fl ow hydrograph. Four<br />

pr<strong>in</strong>ciples which might be used to guide restoration or<br />

enhancement of fl ow for ecologically motivated restoration<br />

schemes based upon common hydrograph characteristics<br />

are described <strong>in</strong> Figure 7.2 (Bunn and Arthr<strong>in</strong>gton,<br />

2002).<br />

From this it can be seen that almost all aspects of <strong>the</strong><br />

fl ow hydrograph have some importance ei<strong>the</strong>r to <strong>in</strong>dividual<br />

species at various stages of <strong>the</strong>ir life cycle, or to <strong>the</strong><br />

determ<strong>in</strong>ation of species assemblages and hence biotic<br />

diversity and abundance more generally. These hydrograph<br />

components may become more or less important <strong>in</strong><br />

<strong>the</strong> context of both <strong>in</strong>ter- and <strong>in</strong>tra-annual streamfl ow<br />

variation. Hypo<strong>the</strong>sised relationships between water<br />

years, hydrograph components and ecological processes<br />

developed for <strong>the</strong> restoration of gravel-bed stream<br />

environments downstream of <strong>the</strong> Sierra Nevada foothills,<br />

California, USA, are illustrated <strong>in</strong> Table 7.1.<br />

Pr<strong>in</strong>ciple 3<br />

access to floodpla<strong>in</strong>s (lateral<br />

connectivity; longitud<strong>in</strong>al<br />

connectivity) essential<br />

Pr<strong>in</strong>ciple 2<br />

regular variation and stable<br />

baseflows determ<strong>in</strong>e life<br />

history patterns<br />

(spawn<strong>in</strong>g/recruitment)<br />

Figure 7.2 Basic pr<strong>in</strong>ciples govern<strong>in</strong>g ecological response to <strong>the</strong> stream hydrograph (Source: modifi ed after Bunn and Arthr<strong>in</strong>gton,<br />

2002)


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 109<br />

Table 7.1 General hypo<strong>the</strong>sised relationships between hydrograph components and ecosystem processes for gravel-bed streams<br />

downstream of <strong>the</strong> Sierra Nevada foothills (Modifi ed from San Joaqu<strong>in</strong> <strong>River</strong> <strong>Restoration</strong> Study Background Report, prepared for<br />

Friant Water Users Authority, L<strong>in</strong>dsay, CA, and Natural Resources Defense Council, San Francisco, CA, (2002). Reproduced with<br />

permission from McBa<strong>in</strong> & Trush, Inc.)<br />

Hydrograph<br />

component<br />

Geomorphic-hydrologic<br />

processes<br />

Snowmelt peak Wetter years: bed mobility,<br />

long duration fl oodpla<strong>in</strong><br />

<strong>in</strong>undation, moderate<br />

channel migration,<br />

groundwater recharge<br />

Normal years: bed<br />

mobility, short duration<br />

fl oodpla<strong>in</strong> <strong>in</strong>undation<br />

Seasonal recession<br />

limb<br />

Summer–fall<br />

basefl ow<br />

W<strong>in</strong>ter–spr<strong>in</strong>g<br />

ra<strong>in</strong>fall run-off<br />

Gradual decrease <strong>in</strong> water<br />

stage, ma<strong>in</strong>ta<strong>in</strong><br />

fl oodpla<strong>in</strong> soil moisture<br />

Wetter years: channel<br />

avulsion, signifi cant<br />

channel migration, bed<br />

scour and deposition,<br />

bed mobility, fl oodpla<strong>in</strong><br />

scour, fl oodpla<strong>in</strong><br />

<strong>in</strong>undation, fi ne<br />

sediment deposition on<br />

fl oodpla<strong>in</strong>s, large<br />

woody debris<br />

recruitment<br />

Normal years: Some<br />

channel migration,<br />

m<strong>in</strong>or bed scour, bed<br />

mobility, fl oodpla<strong>in</strong><br />

<strong>in</strong>undation, some fi ne<br />

sediment deposition on<br />

fl oodpla<strong>in</strong>s<br />

Riparian processes Salmonid life-history processes<br />

Wetter years: riparian<br />

seedl<strong>in</strong>g scour with<strong>in</strong><br />

bankfull channel, riparian<br />

seedl<strong>in</strong>g <strong>in</strong>itiation on<br />

fl oodpla<strong>in</strong>s, discourages<br />

riparian seedl<strong>in</strong>g <strong>in</strong>itiation<br />

with<strong>in</strong> bankfull channel<br />

Normal years: periodic<br />

riparian seedl<strong>in</strong>g <strong>in</strong>itiation<br />

on fl oodpla<strong>in</strong>s<br />

Wetter years: Allow riparian<br />

seedl<strong>in</strong>g establishment on<br />

fl oodpla<strong>in</strong>s<br />

Normal and drier years:<br />

Discourages riparian<br />

seedl<strong>in</strong>g establishment on<br />

fl oodpla<strong>in</strong>s by desiccat<strong>in</strong>g<br />

<strong>the</strong>m, encourage seedl<strong>in</strong>g<br />

establishment with<strong>in</strong><br />

bankfull channel<br />

Encourages late seed<strong>in</strong>g<br />

riparian vegetation<br />

<strong>in</strong>itiation and<br />

establishment with<strong>in</strong><br />

bankfull channel<br />

Wetter years: mature<br />

riparian removal with<strong>in</strong><br />

bankfull channel and<br />

portions of fl oodpla<strong>in</strong>,<br />

scour of seedl<strong>in</strong>gs with<strong>in</strong><br />

bankfull channel, seedbed<br />

creation on fl oodpla<strong>in</strong>s<br />

for new cohort <strong>in</strong>itiation,<br />

microtopography from<br />

fl oodpla<strong>in</strong> scour and fi ne<br />

sediment deposition<br />

Normal years: scour of<br />

seedl<strong>in</strong>gs with<strong>in</strong> bankfull<br />

channel, some fi ne<br />

sediment deposition on<br />

fl oodpla<strong>in</strong>s<br />

Wetter years: Increase juvenile growth rates by<br />

long-term fl oodpla<strong>in</strong> <strong>in</strong>undation, <strong>in</strong>crease<br />

strand<strong>in</strong>g by <strong>in</strong>undat<strong>in</strong>g fl oodpla<strong>in</strong>s,<br />

stimulate outmigration, reduce predation<br />

mortality by reduc<strong>in</strong>g smolt density and<br />

<strong>in</strong>creas<strong>in</strong>g turbidity<br />

Normal years: Increase juvenile growth rates<br />

by short-term fl oodpla<strong>in</strong> <strong>in</strong>undation,<br />

<strong>in</strong>crease strand<strong>in</strong>g by short-term fl oodpla<strong>in</strong><br />

<strong>in</strong>undation, stimulate outmigration, reduce<br />

predation mortality by reduc<strong>in</strong>g smolt<br />

density and <strong>in</strong>creas<strong>in</strong>g turbidity<br />

Drier years: Increase outmigration predation<br />

mortality by <strong>in</strong>creas<strong>in</strong>g density and reduc<strong>in</strong>g<br />

turbidity<br />

Wetter years: Increase outmigration success by<br />

reduc<strong>in</strong>g water temperatures and extend<strong>in</strong>g<br />

outmigration period<br />

Normal years: Increase outmigration success<br />

by reduc<strong>in</strong>g water temperatures and<br />

extend<strong>in</strong>g outmigration period<br />

Drier years: Increase outmigration mortality by<br />

<strong>in</strong>creas<strong>in</strong>g water temperatures and<br />

shorten<strong>in</strong>g outmigration period<br />

Water temperature for over-summer<strong>in</strong>g<br />

juveniles and spr<strong>in</strong>g-run adults, immigration<br />

for fall-run adults<br />

Wetter years: partial loss of cohort due to redd<br />

scour or entombment from deposition,<br />

improve spawn<strong>in</strong>g gravel quality by<br />

scour<strong>in</strong>g/redeposit<strong>in</strong>g bed and transport<strong>in</strong>g<br />

fi ne sediment, mortality by fl ush<strong>in</strong>g fry and<br />

juveniles, mortality by strand<strong>in</strong>g fry and<br />

juveniles on fl oodpla<strong>in</strong>s, reduce growth<br />

dur<strong>in</strong>g periods of high turbidity, reduce<br />

predation dur<strong>in</strong>g periods of high turbidity,<br />

creation and ma<strong>in</strong>tenance of high quality<br />

aquatic habitat<br />

Normal years: improve spawn<strong>in</strong>g gravel<br />

quality by mobilis<strong>in</strong>g bed and transport<strong>in</strong>g<br />

fi ne sediment, low mortality by fl ush<strong>in</strong>g fry<br />

and juveniles, low mortality by strand<strong>in</strong>g fry<br />

and juveniles on fl oodpla<strong>in</strong>s, reduce growth<br />

dur<strong>in</strong>g periods of high turbidity, reduce<br />

predation dur<strong>in</strong>g periods of high turbidity,<br />

ma<strong>in</strong>tenance of high quality aquatic habitat<br />

W<strong>in</strong>ter basefl ows F<strong>in</strong>e sediment transport Increase habitat area <strong>in</strong> natural channel<br />

morphology


110 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

The ‘Hydrograph Component Analysis’ illustrated <strong>in</strong><br />

Table 7.1 enables <strong>the</strong> user to connect <strong>the</strong> geomorphic and<br />

ecological consequences (or responses) to <strong>the</strong> results of<br />

a hydrological analysis. For example, snowmelt peaks<br />

and w<strong>in</strong>ter run-off occurr<strong>in</strong>g <strong>in</strong> wetter years may have<br />

signifi cantly different consequences <strong>in</strong> terms of enhanced<br />

channel change than those occurr<strong>in</strong>g <strong>in</strong> drier years.<br />

Channel changes have associated ecological detriments <strong>in</strong><br />

<strong>the</strong> shorter term, but provide ecological benefi ts <strong>in</strong> <strong>the</strong><br />

longer term. Reduc<strong>in</strong>g hydrological uncerta<strong>in</strong>ties before<br />

restoration (for design purposes) and after restoration (for<br />

monitor<strong>in</strong>g and appraisal purposes) thus requires <strong>the</strong><br />

deployment of techniques that span <strong>the</strong> determ<strong>in</strong>ation of<br />

low fl ows, which characterise <strong>the</strong> occurrence and magnitudes<br />

of high fl ows, and which identify any temporal<br />

changes or trends with respect to <strong>the</strong>se. Subsequently,<br />

l<strong>in</strong>kage of <strong>the</strong>se hydrological parameters to physical (sediment<br />

transport, channel morphological) and ecological<br />

responses may be undertaken. Whereas ‘classical’ hydrology<br />

has focused on determ<strong>in</strong><strong>in</strong>g catchment ‘signatures’<br />

with respect to <strong>the</strong> unit hydrograph (<strong>the</strong> average shape of<br />

storm hydrographs with equal distributions of ra<strong>in</strong>fall),<br />

m<strong>in</strong>imis<strong>in</strong>g uncerta<strong>in</strong>ty <strong>in</strong> restoration applications entails<br />

a more sophisticated paradigm for application. This<br />

requires consideration of hydrograph variability, <strong>the</strong><br />

potential l<strong>in</strong>ks between this and o<strong>the</strong>r properties of <strong>the</strong><br />

catchment network, as well as knowledge of land use<br />

management history. Collectively, it is <strong>the</strong>se which defi ne<br />

a ‘natural fl ow’ as <strong>the</strong> alternative paradigm for hydrological<br />

assessment and river restoration schemes.<br />

While that natural fl ow is now accepted as fundamental<br />

to improved management and study of rivers (Richter et<br />

al., 1996; Poff et al., 1997), it is also becom<strong>in</strong>g clear that<br />

<strong>the</strong> response of organisms to fl ood and drought may be<br />

evidenced over different time scales, and may be fundamentally<br />

different <strong>in</strong> character. Human modifi cations to<br />

<strong>the</strong> fl ow regime may also alter <strong>the</strong> course and viability of<br />

o<strong>the</strong>rwise natural adaptive strategies (Lytle and Poff,<br />

2004). Thus, while general pr<strong>in</strong>ciples or paradigms can be<br />

identifi ed, application to specifi c regions, case studies or<br />

species, may require considerable fi eld calibration, and/or<br />

engender considerable uncerta<strong>in</strong>ty <strong>in</strong> assess<strong>in</strong>g <strong>the</strong> output<br />

of models and scenarios. The follow<strong>in</strong>g sections describe<br />

some of <strong>the</strong> techniques used, and identify areas of uncerta<strong>in</strong>ty<br />

associated with <strong>the</strong>ir application and <strong>in</strong>terpretation<br />

<strong>in</strong> channel restoration designs.<br />

7.3.2 Determ<strong>in</strong><strong>in</strong>g <strong>the</strong> Environmental Flow Regime<br />

for <strong>River</strong> <strong>Restoration</strong>: Some Alternative Frameworks<br />

There is no simple formula or fi gure that can be given for<br />

<strong>the</strong> environmental fl ow requirements or <strong>in</strong>stream fl ow<br />

requirements of rivers. Much depends on <strong>the</strong> desired<br />

future character of <strong>the</strong> river ecosystem under study, which<br />

may be set by legislation or negotiated as a trade-off<br />

between water users. The fl ow allocated to a river may thus<br />

be primarily a matter of social choice, with science provid<strong>in</strong>g<br />

technical support to help determ<strong>in</strong>e <strong>the</strong> river ecosystem<br />

response under various fl ow regimes. Most scientifi c<br />

efforts have been directed to <strong>the</strong> determ<strong>in</strong>ation of m<strong>in</strong>imum<br />

fl ows <strong>in</strong> rivers, particularly dur<strong>in</strong>g dry periods of <strong>the</strong> year<br />

(or <strong>in</strong> drier years). These have been thought to be <strong>the</strong> most<br />

important limit<strong>in</strong>g factor <strong>in</strong> <strong>the</strong> ecological status or health<br />

of <strong>the</strong> river environment. Dur<strong>in</strong>g <strong>the</strong> past 20 years, a range<br />

of methods has been developed to help set environmental<br />

fl ows, each with advantages and disadvantages <strong>in</strong> particular<br />

circumstances of application.<br />

Where fl ow data exist, methods for <strong>the</strong> determ<strong>in</strong>ation<br />

of m<strong>in</strong>imum fl ows have been primarily statistical. Criteria<br />

for method selection <strong>in</strong>clude: <strong>the</strong> type of issue (abstraction,<br />

dam, run-of-river scheme); expertise, time and money<br />

available; and <strong>the</strong> legislative framework with<strong>in</strong> which <strong>the</strong><br />

fl ows must be set. Ungauged catchments give rise to particular<br />

uncerta<strong>in</strong>ties, which may be reduced by construct<strong>in</strong>g<br />

regional regression curves from rivers where data do<br />

exist, supplemented by statistical or geomorphologcal<br />

models of response (see Smakht<strong>in</strong>, 2001 for a comprehensive<br />

review of most methods and applications). Most<br />

recently, means of connect<strong>in</strong>g fl ow to modelled or observed<br />

ecological response have been developed to <strong>in</strong>form hydrological<br />

analyses. There are thus four broad categories of<br />

methods (Acreman and K<strong>in</strong>g, 2003) which may be used<br />

to reduce uncerta<strong>in</strong>ty <strong>in</strong> assess<strong>in</strong>g <strong>the</strong> hydrological basis<br />

for channel restoration designs: look-up tables; desk top<br />

analysis; functional analyses; and habitat modell<strong>in</strong>g. Each<br />

of <strong>the</strong>se methods may <strong>in</strong>volve different degrees of ‘expert’<br />

<strong>in</strong>volvement and may address all or just parts of <strong>the</strong> river<br />

system. Consequently, <strong>the</strong> use of experts and <strong>the</strong> degree<br />

to which methods are holistic are considered as crosscutt<strong>in</strong>g<br />

issues. The basic pr<strong>in</strong>ciples, strengths and weaknesses<br />

of each of <strong>the</strong>se methods are outl<strong>in</strong>ed below.<br />

Look-up Tables<br />

Worldwide, <strong>the</strong> most commonly applied methods to defi ne<br />

target river fl ows are rules-of-thumb based on simple<br />

<strong>in</strong>dices given <strong>in</strong> look-up tables. The most widely employed<br />

<strong>in</strong>dices are purely hydrological but those employ<strong>in</strong>g ecological<br />

data were also developed <strong>in</strong> <strong>the</strong> 1970s. Many early<br />

applications of environmental fl ow sett<strong>in</strong>g focused on<br />

s<strong>in</strong>gle species or s<strong>in</strong>gle issues. For example, much of <strong>the</strong><br />

demand for environmental fl ows <strong>in</strong> North America and<br />

nor<strong>the</strong>rn Europe was from sport fi shermen concerned<br />

about <strong>the</strong> decl<strong>in</strong>e <strong>in</strong> numbers of trout and salmon follow-


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 111<br />

<strong>in</strong>g abstractions and dam operations. Environmental fl ows<br />

were <strong>the</strong>n set to ma<strong>in</strong>ta<strong>in</strong> critical levels of habitat (<strong>in</strong>clud<strong>in</strong>g<br />

sediment, fl ow velocity, depth) for <strong>the</strong>se species. Part<br />

of <strong>the</strong> justifi cation was that <strong>the</strong>se species are very sensitive<br />

to fl ow and if <strong>the</strong> fl ow is appropriate for <strong>the</strong>m, it will be<br />

for o<strong>the</strong>r parts of <strong>the</strong> ecosystem.<br />

Eng<strong>in</strong>eers have traditionally used hydrologicallydefi<br />

ned <strong>in</strong>dices for water management rules to set compensation<br />

fl ows below reservoirs and weirs. Examples are<br />

percentages of <strong>the</strong> mean fl ow or exceedence percentiles<br />

from a fl ow duration curve. (The fl ow duration curve is a<br />

water resources tool that defi nes <strong>the</strong> proportion of time<br />

that a given fl ow is equalled or exceeded). This approach<br />

has been adopted for environmental fl ow sett<strong>in</strong>g to determ<strong>in</strong>e<br />

simple operat<strong>in</strong>g rules for dams or off-take structures<br />

where few or no local ecological data are available.<br />

A hydrological <strong>in</strong>dex is used <strong>in</strong> France, where <strong>the</strong> Freshwater<br />

Fish<strong>in</strong>g Law of 1984 required that residual fl ows <strong>in</strong><br />

bypassed sections of river must be a m<strong>in</strong>imum of 1 – 40 of <strong>the</strong><br />

mean fl ow for exist<strong>in</strong>g schemes and 1 – 10 of <strong>the</strong> mean fl ow<br />

for new schemes (Souchon and Keith, 2001). In Brazil,<br />

fl ows below dams must be at least 80% of m<strong>in</strong>imum<br />

monthly average fl ow (Benetti et al., 2002). In regulat<strong>in</strong>g<br />

abstractions <strong>in</strong> <strong>the</strong> United K<strong>in</strong>gdom, an <strong>in</strong>dex of natural<br />

low fl ow has been employed to defi ne <strong>the</strong> environmental<br />

fl ow. Q 95 (i.e. that fl ow which is equalled or exceeded for<br />

95% of <strong>the</strong> time) is often used. The fi gure of Q 95 was<br />

chosen purely on hydrological grounds. However, <strong>the</strong><br />

implementation of this approach often <strong>in</strong>cludes ecological<br />

<strong>in</strong>formation (Barker and Kirmond, 1998).<br />

The Tennant Method (Tennant, 1976) was developed to<br />

specify m<strong>in</strong>imum fl ows to protect a healthy river environment.<br />

This employed calibration data from hundreds of<br />

rivers <strong>in</strong> <strong>the</strong> mid-Western states of <strong>the</strong> United States. Percentages<br />

of <strong>the</strong> mean annual fl ow are specifi ed that provide<br />

different quality habitat for fi sh, e.g. 10% for poor quality<br />

(survival), 30% for moderate habitat (satisfactory) and<br />

60% for excellent habitat. This approach can be used elsewhere<br />

but <strong>the</strong> exact <strong>in</strong>dices need to be re-calculated for<br />

each new region. The <strong>in</strong>dices are modifi ed where run-off<br />

<strong>in</strong> <strong>the</strong> spr<strong>in</strong>g is important and are widely used <strong>in</strong> plann<strong>in</strong>g<br />

at <strong>the</strong> river bas<strong>in</strong> level.<br />

Mat<strong>the</strong>ws and Bao (1991) concluded that methods<br />

based on proportions of mean fl ow were not suitable for<br />

<strong>the</strong> fl ow regimes of Texan rivers as <strong>the</strong>y often resulted <strong>in</strong><br />

an unrealistically high fl ow. Instead, <strong>the</strong>y devised a method<br />

us<strong>in</strong>g variable percentages of <strong>the</strong> monthly median fl ow,<br />

based on fi sh <strong>in</strong>ventories and known life history requirements,<br />

fl ow frequency distributions and conditions for<br />

special periods and processes (e.g. migration).<br />

The advantage of all look-up approaches is that once<br />

developed, application requires relatively few resources.<br />

However, simple hydrological <strong>in</strong>dices are not readily<br />

transferable between regions without re-calibration. Even<br />

<strong>the</strong>n, <strong>the</strong>y do not take account of site-specifi c conditions.<br />

In particular, adjust<strong>in</strong>g hydrological <strong>in</strong>dices does not<br />

ensure ecological validity without a correspond<strong>in</strong>g<br />

ecological re-assessment, but ecological data may be<br />

much more costly and time consum<strong>in</strong>g to collect. In<br />

general, look-up tables are thus particularly appropriate<br />

for low controversy situations. They also tend to be<br />

precautionary.<br />

Desk Top Analysis<br />

Desk top analyses tend to focus on analysis of exist<strong>in</strong>g<br />

rout<strong>in</strong>e data (such as river fl ows from gaug<strong>in</strong>g stations<br />

and/or fi sh data from regular surveys) although data may<br />

be collected as part of a specifi c project at a particular site<br />

or sites on a river. Some desk top methods are purely<br />

hydrological. For example, Richter et al., (1996, 1997)<br />

developed a hydrological method <strong>in</strong>tended for sett<strong>in</strong>g<br />

benchmark fl ows on rivers, where a natural ecosystem is<br />

<strong>the</strong> primary objective. Development of <strong>the</strong> method relies<br />

upon identifi cation of <strong>the</strong> components of a natural fl ow<br />

regime, <strong>in</strong>dexed by magnitude (of both high and low<br />

fl ows), tim<strong>in</strong>g (<strong>in</strong>dexed by monthly statistics), frequency<br />

(number of events) and duration (<strong>in</strong>dexed by mov<strong>in</strong>g<br />

average m<strong>in</strong>ima and maxima). The method uses gauged or<br />

modelled daily fl ows and a set of 32 <strong>in</strong>dices. A range of<br />

variation of <strong>the</strong> <strong>in</strong>dices may <strong>the</strong>n be set, based upon ±1<br />

standard deviation from <strong>the</strong> mean or between <strong>the</strong> 25th and<br />

75th percentiles. Variability <strong>in</strong> stream fl ow is essential <strong>in</strong><br />

susta<strong>in</strong><strong>in</strong>g ecosystem <strong>in</strong>tegrity (long term ma<strong>in</strong>tenance of<br />

biodiversity and productivity) and resiliency (<strong>the</strong> capacity<br />

to endure natural and human disturbances – Stanford<br />

et al., 1996). This method is <strong>in</strong>tended to defi ne <strong>in</strong>terim<br />

standards, which can be monitored and revised. However,<br />

so far, <strong>the</strong>re has not been enough research to relate <strong>the</strong><br />

fl ow statistics to specifi c elements of <strong>the</strong> ecosystem.<br />

A particular aspect of desk top analysis relates to <strong>the</strong><br />

characterisation of high fl ows <strong>in</strong> rivers, especially <strong>the</strong><br />

identifi cation of group<strong>in</strong>gs of higher fl ow events, ei<strong>the</strong>r<br />

with<strong>in</strong> or between years. The emphasis on higher fl ows is<br />

someth<strong>in</strong>g of a counter to <strong>the</strong> previous concentration of<br />

hydrological analyses on low fl ow requirements. It refl ects<br />

<strong>the</strong> importance of higher magnitude fl ows <strong>in</strong> connect<strong>in</strong>g<br />

<strong>the</strong> various parts of <strong>the</strong> catchment–fl oodpla<strong>in</strong>–channel<br />

system, <strong>in</strong> susta<strong>in</strong><strong>in</strong>g sediment transport through <strong>the</strong><br />

fl uvial system to ma<strong>in</strong>ta<strong>in</strong> channel and fl oodpla<strong>in</strong> structure<br />

(<strong>in</strong>clud<strong>in</strong>g larger-scale riparian woodlands – Gregory et<br />

al., 2003) and <strong>in</strong> servic<strong>in</strong>g particular stages of species’ life<br />

cycles. In this context, fl ow variability or hydrological<br />

disturbance is thus both a potential <strong>in</strong>dicator of land use


112 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

change and an important control on river ecology. Knowledge<br />

of <strong>the</strong> chang<strong>in</strong>g attributes of such disturbance is<br />

important <strong>in</strong> sett<strong>in</strong>g <strong>the</strong> appropriate historical and contemporary<br />

context for stream fl ow restoration (through consideration<br />

of land use and climatic histories), and for<br />

design<strong>in</strong>g-<strong>in</strong> crucial aspects of fl ow variability.<br />

By way of illustration, Archer and Newson (2002) focus<br />

on <strong>the</strong> number and frequency of rises and falls above<br />

selected threshold levels (pulses) expressed as multiples<br />

of <strong>the</strong> median fl ow, toge<strong>the</strong>r with <strong>the</strong> duration of <strong>the</strong> events<br />

obta<strong>in</strong>ed when analys<strong>in</strong>g <strong>the</strong> discharge history of <strong>the</strong> Coalburn<br />

catchment, UK (Figure 7.3). The pulse duration is<br />

<strong>the</strong> time from ris<strong>in</strong>g above <strong>the</strong> threshold to fall<strong>in</strong>g below<br />

<strong>the</strong> same threshold, and <strong>the</strong> number of pulses refl ects <strong>the</strong><br />

magnitude of <strong>the</strong> chosen threshold fl ow. Provided that <strong>the</strong><br />

data series are extensive enough, of suffi cient resolution<br />

(for example, daily mean fl ows are unlikely to be an adequate<br />

basis for analysis when catchment lag is much less<br />

than one day) and are accompanied by o<strong>the</strong>r appropriate<br />

documentary/archival <strong>in</strong>formation, <strong>the</strong> analysis may be<br />

extended to consider periods of land use or climatic<br />

change, as shown <strong>in</strong> Figure 7.4.<br />

The Coalburn is a small upland catchment with an area<br />

of 1.5 km 2 and an altitud<strong>in</strong>al range from 270–330 m. The<br />

natural surface material comprises a cover of blanket peat<br />

(0.5–3 m thick) overly<strong>in</strong>g glacial till up to 5 m <strong>in</strong> thickness.<br />

This was ploughed <strong>in</strong> 1972, result<strong>in</strong>g <strong>in</strong> a dra<strong>in</strong>age density<br />

60 times greater than <strong>the</strong> orig<strong>in</strong>al stream network. In <strong>the</strong><br />

spr<strong>in</strong>g of 1973, 90% of <strong>the</strong> catchment was planted with<br />

Sitka spruce and, s<strong>in</strong>ce <strong>the</strong>n, growth rates have been variable,<br />

reach<strong>in</strong>g 1 m height <strong>in</strong> 1978 and 7–12 m <strong>in</strong> 1996, by<br />

which time some 60% of <strong>the</strong> catchment had reached<br />

canopy closure. With this knowledge of land use management,<br />

<strong>the</strong> hydrological data can be analysed as shown <strong>in</strong><br />

Discharge<br />

1 2<br />

1<br />

1<br />

2<br />

Time<br />

Figure 7.4 to reveal phases of hydrological response to <strong>the</strong><br />

land use change. Thus, immediately follow<strong>in</strong>g plough<strong>in</strong>g<br />

and plant<strong>in</strong>g, pulse numbers and total pulse duration<br />

<strong>in</strong>crease, and subsequently decl<strong>in</strong>e with vegetation maturity,<br />

whereas <strong>the</strong> average duration of <strong>in</strong>dividual events<br />

<strong>in</strong>creases as vegetation matures. These patterns <strong>in</strong> hydrological<br />

response are most clearly shown when <strong>the</strong> analysis<br />

is applied to events defi ned by thresholds of 2–6 times <strong>the</strong><br />

median fl ow. Importantly, <strong>the</strong> method fur<strong>the</strong>r demonstrates<br />

that, even where peak and time-to-peak of <strong>the</strong> unit hydrograph<br />

are similar <strong>in</strong> pre- and post-disturbance situations,<br />

<strong>the</strong>re are differences <strong>in</strong> pulse numbers and pulse magnitudes,<br />

which are thought to help determ<strong>in</strong>e habitat potential<br />

and ecological response. Not surpris<strong>in</strong>gly given<br />

its emphasis on higher magnitude threshold events, <strong>the</strong><br />

method is least satisfactory with respect to disclos<strong>in</strong>g<br />

behaviour of low fl ows, and this limitation is important <strong>in</strong><br />

relation to use of <strong>the</strong> technique to ‘set’ regulated low fl ows<br />

as described above. In such cases, it may, however, be<br />

supplemented by o<strong>the</strong>r procedures and <strong>the</strong> method should<br />

be thought of as complement<strong>in</strong>g traditional approaches so<br />

as to <strong>in</strong>clude hydrological variability as a key ecological<br />

determ<strong>in</strong>ant.<br />

Ano<strong>the</strong>r area of research that may be considered under<br />

<strong>the</strong> head<strong>in</strong>g of desk top analysis (but which underp<strong>in</strong>s<br />

o<strong>the</strong>r methodologies described below) is <strong>the</strong> prediction of<br />

hydrograph characteristics from catchment channel<br />

network properties. Here, <strong>the</strong> focus of attention has been<br />

on <strong>the</strong> effects of network scale and <strong>the</strong> relative tim<strong>in</strong>g<br />

of <strong>the</strong> hillslope hydrograph and channel water rout<strong>in</strong>g.<br />

Because <strong>the</strong> density of channels <strong>in</strong> a river network refl ects<br />

climate and land use, <strong>the</strong>re should be a network response<br />

(transformation) to any changes <strong>in</strong> <strong>the</strong>se driv<strong>in</strong>g variables<br />

(Kirkby, 1993). The goal has been to <strong>in</strong>corporate network<br />

2 3<br />

3 × Median<br />

2 × Median<br />

Median<br />

3<br />

4 5<br />

Figure 7.3 Defi nition diagram of pulses above selected thresholds and pulse duration (between arrows) <strong>in</strong> hydrological time series<br />

(Repr<strong>in</strong>ted from D. Archer et al. (2002), Journal of Hydrology 268, 244–258, with permission from Elsevier.)


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 113<br />

Annual number of pulses<br />

Total duration (hrs) above threshold<br />

Average duration of flow above<br />

80<br />

70<br />

60<br />

50<br />

40<br />

20<br />

10<br />

0<br />

5000<br />

4500<br />

4000<br />

3500<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

threshold (hrs) 30<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

>0.5M<br />

>M<br />

>2M<br />

>3M<br />

>4M<br />

>0.5M<br />

>M<br />

>2M<br />

>3M<br />

>4M<br />

>0.5M<br />

>M<br />

>2M<br />

>3M<br />

>4M<br />

>5M<br />

>5M<br />

>5M<br />

>6M<br />

>7M<br />

>8M<br />

>10M<br />

>15M<br />

(a)<br />

(b)<br />

(c)<br />

>20M<br />

Multiples of median flow<br />

>6M<br />

>7M<br />

>8M<br />

>10M<br />

>15M<br />

>20M<br />

>30M<br />

>40M<br />

>30M<br />

Multiples of median flow<br />

>6M<br />

>7M<br />

>8M<br />

>10M<br />

>15M<br />

>20M<br />

>30M<br />

Multiples of median flow<br />

>50M<br />

>60M<br />

1967-99<br />

1967-71<br />

1974-82<br />

1983-90<br />

1992-99<br />

>80M<br />

>40M<br />

>50M<br />

>60M<br />

>80M<br />

>40M<br />

>50M<br />

>60M<br />

>80M<br />

>100M<br />

1967-99<br />

1967-71<br />

1974-82<br />

1983-90<br />

1992-99<br />

>100M<br />

1967-99<br />

1967-71<br />

1974-82<br />

1983-90<br />

1992-99<br />

Figure 7.4 Pulse number (a), total pulse duration (b) and average pulse duration (c) for <strong>the</strong> Coalburn catchment over <strong>the</strong> full range<br />

of fl ows <strong>in</strong> pre- and post-dra<strong>in</strong>age and plant<strong>in</strong>g periods (Repr<strong>in</strong>ted from D. Archer et al. (2002), Journal of Hydrology 268, 244–258,<br />

with permission from Elsevier.)<br />

>100M


114 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

geometry <strong>in</strong>to a catchment hydrograph forecast<strong>in</strong>g model<br />

and this has given rise to <strong>the</strong> Geomorphological Unit<br />

Hydrograph (GUH). This defi nes <strong>the</strong> average shape of<br />

storm hydrographs with equal distributions of ra<strong>in</strong>fall as<br />

a function of bifurcation ratio, bas<strong>in</strong> area ratio, stream<br />

length ratio, watershed order and a mean wait<strong>in</strong>g time for<br />

each order (Rodriguez-Iturbe and Valdes, 1979):<br />

and<br />

peak discharge = 1.31/ L ΩRL 0.43<br />

time to peak = 0.44 L Ω RB 0.55 RA −0.55 RL −0.38<br />

where RB = bifurcation ratio; RA = bas<strong>in</strong> area ratio; RL =<br />

stream length ratio; and bas<strong>in</strong> order = Ω.<br />

The GUH may be criticised because of its dependence<br />

on network properties that are apparently <strong>in</strong>sensitive to<br />

changes and its neglect of some aspects describ<strong>in</strong>g network<br />

shape, topology and topography. More recent developments<br />

of this approach are directed to <strong>in</strong>corporat<strong>in</strong>g<br />

measures of channel network width and network l<strong>in</strong>k concentration<br />

functions (for detailed review, see Mousouridis,<br />

2001).<br />

O<strong>the</strong>r research has also related river fl ow directly to<br />

ecological data, such as population numbers or <strong>in</strong>dices<br />

of community structure calculated from species lists.<br />

However, it is diffi cult to derive biotic <strong>in</strong>dices that are only<br />

sensitive to fl ow and not to o<strong>the</strong>r factors (e.g. habitat<br />

structure or water quality). At <strong>the</strong> m<strong>in</strong>imum, biotic <strong>in</strong>dices<br />

designed for water quality monitor<strong>in</strong>g purposes should be<br />

used with extreme caution (Armitage and Petts, 1992).<br />

Generally, such data are scarce, and <strong>in</strong>terpretation is complicated<br />

where methods of collection encompass po<strong>in</strong>tspecifi<br />

c measurements obta<strong>in</strong>ed at different times, and<br />

spatially-distributed methods pick<strong>in</strong>g-out more persistent<br />

patterns refl ect<strong>in</strong>g longer term aspects of land use and<br />

water quality (Dakova et al., 2000). In addition, it is not<br />

always clear which fl ow variables to choose to represent<br />

different aspects of <strong>the</strong> fl ow regime which are of<br />

most ecological relevance. Studies which directly relate<br />

‘responses’ such as macro<strong>in</strong>vertebrate abundance to<br />

hydro-climatological and sediment variables confi rm <strong>the</strong><br />

importance of <strong>in</strong>termediate levels of disturbance as underp<strong>in</strong>n<strong>in</strong>g<br />

greatest habitat diversity, and thus imply that<br />

modell<strong>in</strong>g must <strong>in</strong>corporate more than one aspect of<br />

hydrological diversity (Reice et al., 1990).<br />

The lack of complementary hydrological and biological<br />

data is often a limit<strong>in</strong>g factor, and sometimes rout<strong>in</strong>ely<br />

collected data ga<strong>the</strong>red for o<strong>the</strong>r purposes turn out to be<br />

unsuitable. In addition, time series may not be <strong>in</strong>dependent<br />

(which can violate assumptions of classical statistical<br />

techniques) while <strong>the</strong> representation and characterisation<br />

of disturbances (‘extremes’) necessitates longer, higherquality<br />

time series. For example, Clausen and Biggs (2000)<br />

exam<strong>in</strong>ed 35 fl ow variables us<strong>in</strong>g daily mean fl ows for a<br />

seven-year record common to 62 perennial rivers <strong>in</strong> New<br />

Zealand. Based upon a covariance analysis among <strong>the</strong> sites<br />

through a pr<strong>in</strong>cipal components analysis, <strong>the</strong> 35 variables<br />

could be collapsed <strong>in</strong>to four variable group<strong>in</strong>gs relat<strong>in</strong>g<br />

to: size of river; overall variability of <strong>the</strong> fl ow; volume of<br />

high fl ow; and frequency of high fl ow. Signifi cantly, <strong>the</strong><br />

statistical properties (particularly <strong>in</strong>tra-annual variability)<br />

varied between groups, necessitat<strong>in</strong>g <strong>the</strong> use of a suite of<br />

different variables from each group to adequately represent<br />

<strong>the</strong> facets of fl ow of most ecological relevance. This illustrates<br />

<strong>the</strong> requirements for high quality data and, too, <strong>the</strong><br />

uncerta<strong>in</strong>ties of <strong>in</strong>terpretation and application.<br />

A method recently developed <strong>in</strong> <strong>the</strong> United K<strong>in</strong>gdom<br />

<strong>in</strong>volv<strong>in</strong>g ecological data is <strong>the</strong> Lotic Invertebrate Index<br />

for Flow Evaluation (LIFE; Extence et al., 1999). The<br />

LIFE score is based on <strong>the</strong> abundance and sensitivity to<br />

water velocity of different taxa, collected <strong>in</strong> rout<strong>in</strong>e macro<strong>in</strong>vertebrate<br />

monitor<strong>in</strong>g data. Mov<strong>in</strong>g averages of preced<strong>in</strong>g<br />

fl ow have shown good relationships with LIFE scores<br />

over a range of sites (Figure 7.5), but it is as yet uncerta<strong>in</strong><br />

as to how <strong>the</strong> approach can be used to manage river fl ows.<br />

Never<strong>the</strong>less, <strong>the</strong> pr<strong>in</strong>ciple is believed to be sound and<br />

LIFE has <strong>the</strong> major advantage of utilis<strong>in</strong>g <strong>the</strong> data collected<br />

by exist<strong>in</strong>g bio-monitor<strong>in</strong>g programmes.<br />

Functional Analysis Methods – <strong>the</strong> Build<strong>in</strong>g<br />

Block Methodology<br />

The third group of methods to determ<strong>in</strong>e fl ow requirements<br />

for restoration and ecological purposes <strong>in</strong>cludes<br />

those that build an understand<strong>in</strong>g of <strong>the</strong> functional l<strong>in</strong>ks<br />

between hydrology and ecology <strong>in</strong> <strong>the</strong> entire river system.<br />

These methods take a broad view and comb<strong>in</strong>e hydrological<br />

analysis, hydraulic rat<strong>in</strong>g <strong>in</strong>formation (to estimate<br />

sediment transport and channel capacity) and many aspects<br />

of <strong>the</strong> river ecosystem. Perhaps <strong>the</strong> best known is <strong>the</strong><br />

Build<strong>in</strong>g Block Methodology (BBM) developed <strong>in</strong> South<br />

Africa (Rowntree and Wadeson, 1998; K<strong>in</strong>g et al., 2000).<br />

The basic premise of <strong>the</strong> BBM is that river<strong>in</strong>e species are<br />

reliant on three basic elements (build<strong>in</strong>g blocks) of <strong>the</strong><br />

fl ow regime, each of which has a particular ecological and<br />

geomorphological signifi cance: low fl ows, freshets and<br />

fl oods. An acceptable fl ow regime for ecosystem ma<strong>in</strong>tenance<br />

can thus be constructed by comb<strong>in</strong><strong>in</strong>g <strong>the</strong>se build<strong>in</strong>g<br />

blocks. In a typical exercise to determ<strong>in</strong>e <strong>in</strong>stream fl ow<br />

requirements through <strong>the</strong> build<strong>in</strong>g block method, fi ve<br />

stages may be <strong>in</strong>volved (Rowntree and Wadeson, 1998): a<br />

general assessment of catchment condition to determ<strong>in</strong>e


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 115<br />

Discharge (m³/s)<br />

10.00<br />

1.00<br />

0.10<br />

0.01<br />

0.00<br />

6.0<br />

86 87 88 88 89 90 91 92 93 94 95 96 97 98 99 00<br />

Year<br />

Flow<br />

Figure 7.5 Example <strong>River</strong> Flow (logarithmic scale) and LIFE Score time series<br />

potential for geomorphological change <strong>in</strong> a river; an evaluation<br />

of river network characteristics to help identify representative<br />

river reaches to determ<strong>in</strong>e <strong>the</strong> <strong>in</strong>stream fl ow<br />

requirements; an assessment of <strong>the</strong> hydraulic diversity at<br />

various stages for each of <strong>the</strong> morphological units present<br />

<strong>in</strong> <strong>the</strong>se reaches (see below for discussion of biotopes); an<br />

assessment of <strong>the</strong> magnitude and importance of freshet<br />

and fl oods; and an assessment of any <strong>in</strong>ter-bas<strong>in</strong> water<br />

transfers.<br />

The BBM thus revolves around a team of physical scientists<br />

(e.g. hydrologists and geomorphologists) and biological<br />

scientists (e.g. botanists and fi sh biologists) who<br />

follow a series of structured analyses to come to a consensus<br />

on <strong>the</strong> build<strong>in</strong>g blocks of <strong>the</strong> fl ow regime. The BBM<br />

has a detailed manual for implementation (K<strong>in</strong>g et al.,<br />

2000); is rout<strong>in</strong>ely used <strong>in</strong> South Africa to comply with<br />

<strong>the</strong> 1998 Water Act; has been applied <strong>in</strong> Australia (Arth<strong>in</strong>gton<br />

and Long 1997, Arth<strong>in</strong>gton and Lloyd, 1998); and is<br />

be<strong>in</strong>g trialled <strong>in</strong> <strong>the</strong> United States.<br />

To overcome <strong>the</strong> diffi culties <strong>in</strong> relat<strong>in</strong>g changes <strong>in</strong> <strong>the</strong><br />

fl ow regime directly to <strong>the</strong> response of multiple species<br />

and communities, approaches have been developed that<br />

use habitat for target species as an <strong>in</strong>termediate step. Of<br />

those environmental conditions required by an <strong>in</strong>dividual<br />

animal liv<strong>in</strong>g <strong>in</strong> a river, it is <strong>the</strong> physical aspects that are<br />

most heavily impacted by changes to <strong>the</strong> fl ow regime.<br />

Habitat models seek to l<strong>in</strong>k data on <strong>the</strong> physical conditions<br />

(such as water depths and velocities) <strong>in</strong> rivers at different<br />

fl ows (ei<strong>the</strong>r measured data or derived from computer<br />

models) with data on those physical conditions which key<br />

animal or plant species (or <strong>the</strong>ir <strong>in</strong>dividual developmental<br />

stages) require. Once functional relationships between<br />

physical habitat and ecology have been defi ned, <strong>the</strong>y are<br />

l<strong>in</strong>ked to fl ow scenarios <strong>in</strong> river restoration and design (see<br />

Section 7.3 for fur<strong>the</strong>r details on habitat modell<strong>in</strong>g).<br />

Life Score<br />

8.0<br />

7.5<br />

7.0<br />

6.5<br />

LIFE Score<br />

The Holistic Approach<br />

Where <strong>in</strong>tervention seeks fl ow remediation or restoration,<br />

<strong>the</strong>re is an implicitly holistic approach <strong>in</strong>sofar as all elements<br />

of <strong>the</strong> river ecosystem are likely to be supported.<br />

However, more and more methods now adopt an overtly<br />

holistic assessment of <strong>the</strong> whole ecosystem, such as<br />

associated wetlands, groundwater and estuaries; all<br />

species that are sensitive to fl ow (<strong>in</strong>vertebrates, plants and<br />

animals); <strong>the</strong> human context; and all aspects of <strong>the</strong> hydrological<br />

regime, <strong>in</strong>clud<strong>in</strong>g fl oods, droughts, and water<br />

quality. A fundamental pr<strong>in</strong>ciple is to ma<strong>in</strong>ta<strong>in</strong> natural<br />

variability of fl ows and to allocate fl ow based upon likely<br />

habitat impact. This ecosystem approach is especially<br />

important <strong>in</strong> <strong>the</strong> management and restoration of dryland<br />

environments, where <strong>in</strong>herent variability of hydrological<br />

‘extremes’ is a limit<strong>in</strong>g factor ecologically (Thoms and<br />

Sheldon, 2002). Ecosystem approaches <strong>in</strong>herently allow<br />

for complexity and diversity, and may also capitalise on<br />

features such as persistence and evolution – ecological<br />

systems may have self-organis<strong>in</strong>g or self-design<strong>in</strong>g<br />

attributes (Bergen et al., 2001).<br />

Generally, holistic approaches make use of teams of<br />

experts and may also <strong>in</strong>volve <strong>the</strong> participation of stakeholders,<br />

thus extend<strong>in</strong>g holism beyond scientifi c issues.<br />

Where methods have <strong>the</strong> characteristic of be<strong>in</strong>g holistic<br />

<strong>the</strong>y clearly have <strong>the</strong> advantage of cover<strong>in</strong>g <strong>the</strong> whole<br />

hydrological–ecological–stakeholder system. The disadvantage<br />

is that it is diffi cult to identify and to collect <strong>the</strong><br />

relevant data which capture such diverse perspectives.<br />

7.3.3 Choice of Method<br />

Selection of <strong>the</strong> most appropriate method for l<strong>in</strong>k<strong>in</strong>g<br />

hydrological criteria to river restoration designs, where a


116 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Table 7.2 Some advantages and disadvantages of different methods and characteristics of sett<strong>in</strong>g environmental fl ows<br />

Method type Sub-type Advantages Disadvantages and Uncerta<strong>in</strong>ties<br />

Look-up table Hydrological Cheap, rapid to use once calculated Not site-specifi c. Hydrological <strong>in</strong>dices not<br />

valid ecologically<br />

Ecological Ecological <strong>in</strong>dices need region-specifi c data<br />

to be calculated<br />

Desk top Hydrological Site specifi c<br />

Limited new data collection<br />

Long time series required<br />

Hydraulic No explicit use of ecological data<br />

Ecological Ecological data time consum<strong>in</strong>g to collect<br />

Functional analysis Flexible, robust, more focused on<br />

whole ecosystem<br />

range (or set) of methods is available, is considered briefl y<br />

here. Some of <strong>the</strong> advantages and disadvantages which<br />

underlie <strong>the</strong> uncerta<strong>in</strong>ties <strong>in</strong> us<strong>in</strong>g different approaches are<br />

summarised <strong>in</strong> Table 7.2.<br />

Broadly, Table 7.2 illustrates two trade-offs: that between<br />

local sensitivity or regional coverage, and that between<br />

data availability and complexity or simplicity of data <strong>in</strong>terpretation.<br />

Mov<strong>in</strong>g through <strong>the</strong> table, <strong>the</strong> more <strong>in</strong>tegrated<br />

and hence more complex methods are characterised by a<br />

greater degree of connection between physical and biological<br />

parameters. The key element of this <strong>in</strong>tegration is to<br />

describe or model <strong>the</strong> effects of fl ows on habitat structure<br />

and function. In turn, this necessitates <strong>the</strong> addition of<br />

hydraulic and hydrodynamic considerations to supplement<br />

<strong>the</strong> hydrological foundations described above.<br />

7.4 HYDRAULIC ASPECTS AND UNCERTAINTY<br />

IN CHANNEL RESTORATION<br />

7.4.1 Connect<strong>in</strong>g Flows, Sediments and Ecological<br />

Response: <strong>the</strong> Physical Habitat and Eco-Hydraulics<br />

Over <strong>the</strong> past decade, <strong>the</strong>re has been a trend towards<br />

approach<strong>in</strong>g river habitat assessment and rehabilitation<br />

design us<strong>in</strong>g comb<strong>in</strong>ations of fi eld survey and predictive<br />

hydraulic models (Chapter 5). ‘Eco-hydraulics’ (Leclerc,<br />

2002) is now a commonplace term <strong>in</strong> both <strong>the</strong> academic<br />

and practitioner literature, and <strong>the</strong> l<strong>in</strong>kage between physical,<br />

chemical and biotic components of <strong>the</strong> river environment<br />

is central <strong>in</strong> efforts to restore and ma<strong>in</strong>ta<strong>in</strong> ecological<br />

habitat and function (Kemp et al., 2000). To realise <strong>the</strong>se<br />

benefi ts, however, cost-effective (but scientifi cally sound)<br />

means of comb<strong>in</strong><strong>in</strong>g traditional fi eld monitor<strong>in</strong>g and<br />

Expensive to collect all relevant data and to<br />

employ wide range of experts. Consensus<br />

of experts may not be achieved.<br />

Habitat modell<strong>in</strong>g Replicable, predictive Expensive to collect hydraulic and ecological<br />

data<br />

survey with emerg<strong>in</strong>g modell<strong>in</strong>g and design-support<br />

approaches are required. There rema<strong>in</strong>s much work to be<br />

done both to encourage multi-discipl<strong>in</strong>ary co-operation<br />

and to ‘<strong>in</strong>vent’ new trans-discipl<strong>in</strong>ary areas of research<br />

and practical expertise (Janauer, 2000).<br />

Physical habitat can be defi ned as a set of physical<br />

conditions that can be measured and compared to <strong>the</strong><br />

conditions that may be suitable for specifi c species or<br />

<strong>in</strong>dividuals at a particular stage of <strong>the</strong>ir life cycle. The<br />

fundamental aspects of <strong>the</strong> eco-hydraulic approach to river<br />

characterisation and restoration are summarised <strong>in</strong> Figure<br />

7.6. This is based upon <strong>the</strong> idea that species assemblages<br />

and/or abundances are organised to refl ect physical conditions<br />

at a range of scales: catchment conditions (such as<br />

slope, geology and ra<strong>in</strong>fall regime) determ<strong>in</strong>e at <strong>the</strong><br />

most fundamental level <strong>the</strong> k<strong>in</strong>d of species which may be<br />

present; at <strong>the</strong> reach scale, assemblages or abundances<br />

from with<strong>in</strong> this broader range are most likely to be<br />

observed depend<strong>in</strong>g on fl ow regime; at <strong>the</strong> sub-reach<br />

scale, particular species or <strong>in</strong>dividuals at particular stages<br />

of <strong>the</strong>ir life cycle, are found <strong>in</strong> niches with particular fl ow<br />

and sedimentological conditions.<br />

While <strong>the</strong> catchment scale conditions <strong>the</strong> range of<br />

habitat potential, at <strong>the</strong> smaller scales physical habitat is<br />

commonly defi ned by comb<strong>in</strong>ations of depth, velocity,<br />

substrate and distance to cover. In this respect, physical<br />

habitat quality is normally considered <strong>in</strong>dependently of<br />

water quality issues. It may be approached (at its simplest)<br />

from visual survey, through statistical models based upon<br />

measured associations, and via complex determ<strong>in</strong>istic<br />

modell<strong>in</strong>g, <strong>in</strong> which <strong>the</strong> physical habitat is fi rst modelled<br />

and <strong>the</strong>n l<strong>in</strong>ked to ecological response. Although <strong>the</strong><br />

potential of eco-hydraulic modell<strong>in</strong>g has been generally


DRAINAGE<br />

BASIN<br />

10 6 -10 5 years<br />

Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 117<br />

10 4 -10 3 m<br />

10 3 -10 2 m<br />

FLOODPLAIN<br />

10 4 -10 3 years<br />

10 2 -10 1 m<br />

REACH<br />

10 2 -10 1 years<br />

HABITAT<br />

10 1 -10 0 years<br />

10 –1 m<br />

Sand silt<br />

over cobbles<br />

Gravel<br />

MICROHABITAT<br />

10 0 -10 –1 years<br />

Aquatic and<br />

semi-aquatic<br />

vegetation<br />

Leaf and stick<br />

detritus <strong>in</strong><br />

marg<strong>in</strong><br />

Figure 7.6 Spatial scale and distribution of stream habitats based upon fl ow–sediment <strong>in</strong>teractions (Source: Naiman et al., 1992<br />

after Frissell et al., 1986)<br />

Fall<br />

Rapid<br />

Marg<strong>in</strong>al<br />

Deadwater<br />

Chute<br />

acknowledged positively, some signifi cant areas of uncerta<strong>in</strong>ty<br />

rema<strong>in</strong>. A fundamental requirement is <strong>the</strong> ability<br />

to demonstrate (ra<strong>the</strong>r than assume) a l<strong>in</strong>kage between<br />

measured parameters delimit<strong>in</strong>g physical habitat and<br />

<strong>the</strong> species occurrence and abundances (‘responses’ or<br />

‘tolerances’) which <strong>the</strong>se are supposed to condition.<br />

7.4.2 Survey and Visual Methods: Biotopes and<br />

Functional Habitats<br />

At present, hydraulic performance and habitat are essentially<br />

approached separately. Two alternatives are available:<br />

‘bottom up’, <strong>in</strong> which <strong>in</strong>-stream habitat units are<br />

PHYSICAL BIOTOPES FUNCTIONAL (MESO-) HABITATS<br />

Riffle<br />

Boll<br />

Run<br />

glide<br />

Pool<br />

Float<strong>in</strong>g<br />

leaves<br />

Submerged<br />

plants<br />

(slow)<br />

Cobbles/<br />

pobbles<br />

Rock Woody<br />

debris<br />

Tree roots/<br />

overhang<strong>in</strong>g<br />

vegetation<br />

(fast)<br />

Gravel Emergent<br />

plants<br />

Sand<br />

Silt<br />

Marg<strong>in</strong>al<br />

plants<br />

Figure 7.7 Representation of physical biotopes and functional habitats <strong>in</strong> a stream sub-reach (Reproduced from M. D. Newson and<br />

C. L. Newson (2000) ‘Geomorphology, ecology and river channel habitat: mesoscale approaches to bas<strong>in</strong>-scale challenges’ Progress<br />

<strong>in</strong> Physical Geography 24, 195–217, with <strong>the</strong> permission of Edward Arnold (Publishers) Ltd.)<br />

<strong>in</strong>ferred from a knowledge of hydraulic conditions that<br />

defi ne physical biotopes (channel features characterised<br />

through hydraulic measurement or visual survey of surface<br />

fl ow type; Padmore, 1997); and ‘top down’, <strong>in</strong> which<br />

functional or meso- habitats are <strong>in</strong>ferred from analysis of<br />

biological communities associated with substrate and vegetation<br />

characteristics (Kemp et al., 1999). A common<br />

strategy is to visually classify velocity–depth comb<strong>in</strong>ations<br />

(i.e. <strong>the</strong> biotopes) which might <strong>the</strong>n be associated<br />

with functional habitats, or discrete species assemblages<br />

as <strong>in</strong>dicated <strong>in</strong> Figure 7.7 (Newson and Newson, 2000).<br />

Biotopes are easily recognised from fi eld survey<br />

without costly fi eld measurement and, when l<strong>in</strong>ked with


118 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

knowledge of habitat suitability, allow some <strong>in</strong>sight <strong>in</strong>to<br />

likely biological function. Despite early optimism with<br />

respect to this approach, relat<strong>in</strong>g biotopes to mean<strong>in</strong>gful<br />

biotic function appears more complicated than fi rst<br />

assumed (Kemp et al., 2000; Leclerc, 2002). Demonstrable<br />

relationships between biotopes and functional habitats<br />

rema<strong>in</strong> scarce. While <strong>the</strong> survey or biotope method is<br />

meant to circumvent issues relat<strong>in</strong>g to s<strong>in</strong>gle-species preference<br />

curves or s<strong>in</strong>gle measures of channel environment<br />

(s<strong>in</strong>ce <strong>the</strong> biotope is an ‘<strong>in</strong>tegrated’ habitat), major uncerta<strong>in</strong>ties<br />

<strong>in</strong> improv<strong>in</strong>g l<strong>in</strong>ks between biotopes (from simple<br />

survey or modell<strong>in</strong>g) and functional habitats rema<strong>in</strong><br />

(Newson and Newson, 2000). In part, <strong>the</strong> scarcity of data<br />

refl ects <strong>the</strong> cost and practical limitations associated with<br />

fi ne-scale fi eld survey (Newson et al., 1998) and diffi culties<br />

<strong>in</strong> monitor<strong>in</strong>g highly dynamic, stage-dependent fl ow<br />

characteristics. There is also <strong>the</strong> question of <strong>the</strong> appropriate<br />

range of supposed habitat determ<strong>in</strong>ants to associate<br />

with particular species or species clusters, at differ<strong>in</strong>g<br />

stages of <strong>the</strong>ir life cycles, and which may vary greatly<br />

between species found even <strong>in</strong> close proximity. Key questions<br />

which belie <strong>the</strong> uncerta<strong>in</strong>ty <strong>in</strong> an o<strong>the</strong>rwise appeal<strong>in</strong>g<br />

approach are: how consistent are our observations?<br />

How stable are biotopes with stage? How do biotopes<br />

relate to channel morphology and biology?<br />

Biotope mapp<strong>in</strong>g has been supported by ecological<br />

observations that species patch<strong>in</strong>ess (superimposed on a<br />

general downstream cont<strong>in</strong>uum) does occur, and that this<br />

may fur<strong>the</strong>r be related to discharge exceedence values<br />

(Newson and Newson, 2000). This suggests that <strong>the</strong>re<br />

should be a l<strong>in</strong>k with similar patch<strong>in</strong>ess of functional<br />

habitat, if this, too, can be defi ned. Functional habits may<br />

be discrim<strong>in</strong>ated on <strong>the</strong> basis of comb<strong>in</strong>ed fl ow <strong>in</strong>dices<br />

such as <strong>the</strong> Froude number. (The Froude number is <strong>the</strong><br />

ratio of depth to <strong>the</strong> square root of depth multiplied by<br />

gravitational acceleration). From a survey of 32 sites <strong>in</strong><br />

eastern England, Kemp et al. (2000) suggest that functional<br />

habits fall <strong>in</strong>to two groups: lowest Froude number<br />

classes, and those with Froude number above 0.5. Insofar<br />

as <strong>the</strong> Froude number itself is assumed to refl ect biotopes<br />

derived from measurement or observation, <strong>the</strong>n a connection<br />

between biotopes and functional habitats should exist,<br />

but <strong>the</strong> ecological signifi cance of this l<strong>in</strong>k is uncerta<strong>in</strong> <strong>in</strong><br />

<strong>the</strong> absence of simultaneous, po<strong>in</strong>t-by-po<strong>in</strong>t measures of<br />

physical and ecological <strong>in</strong>dicators, and because of feedbacks<br />

or <strong>in</strong>terdependence between <strong>the</strong> apparently <strong>in</strong>dependent<br />

classifi cation variables. For example, vegetation<br />

growth affects velocities and sedimentation, and hence<br />

helps determ<strong>in</strong>e <strong>the</strong> local Froude number, which is supposedly<br />

<strong>in</strong>dependent. A recent <strong>in</strong>vestigation by Clifford<br />

et al. (2006) identifi es <strong>the</strong> need for more consistent use of<br />

bitope and habitat defi nitions, and improved experimental<br />

design when assess<strong>in</strong>g <strong>the</strong> l<strong>in</strong>kage between fl ows, habitats<br />

and ecology. Use of <strong>the</strong> Froude number may also be questioned<br />

because it may obscure differences between very<br />

different fl ow–depth comb<strong>in</strong>ations.<br />

Figure 7.8 illustrates ano<strong>the</strong>r basic area of uncerta<strong>in</strong>ty<br />

<strong>in</strong> <strong>the</strong> biotope and physical habitat approach, but also<br />

perhaps one way <strong>in</strong> which this uncerta<strong>in</strong>ty might be<br />

managed or, <strong>in</strong> future, reduced. The fi gures show reaches<br />

of <strong>the</strong> <strong>River</strong> Cole, near Birm<strong>in</strong>gham, UK, and <strong>the</strong> <strong>River</strong><br />

Tern, Shropshire, UK, which have been ‘classifi ed’ <strong>in</strong>to<br />

zones of statistically similar velocity and/or depth behaviour<br />

at various fl ow stages (Emery et al., 2003). Both<br />

rivers have a well-marked riffl e-pool bedform sequence (a<br />

common biotope), but <strong>the</strong> Cole is a less s<strong>in</strong>uous channel,<br />

with more regular bank morphology and less bankside tree<br />

growth. The approach uses cluster analysis (a hierarchical<br />

statistical method of association) as a reproducible means<br />

of assess<strong>in</strong>g both <strong>the</strong> degree, location and persistence of<br />

cluster<strong>in</strong>g or patch<strong>in</strong>ess of physical habitat, which might<br />

support more objective identifi cation of biotopes.<br />

Cluster analysis is an agglomerative process to identify<br />

homogeneous group<strong>in</strong>gs (patches) based upon selected<br />

characteristics. In this case, standardised fi eld velocity<br />

(measured velocity scaled accord<strong>in</strong>g to <strong>the</strong> mean and variance<br />

of <strong>the</strong> total measured distribution) was used. Initially,<br />

all velocity observations were considered as separate, but<br />

were <strong>the</strong>n successively comb<strong>in</strong>ed accord<strong>in</strong>g to ‘distance’<br />

(i.e. differences) from neighbours and from emerg<strong>in</strong>g<br />

clusters. The next stage was to assess cluster number <strong>in</strong><br />

relation to morphology, spatial coverage and stability.<br />

ANOVA (analysis of variance test<strong>in</strong>g) was used to test<br />

differ<strong>in</strong>g cluster numbers for statistically signifi cant differences<br />

at each fl ow stage and to assess <strong>the</strong>ir coverage<br />

and coherence as fl ow stage varied. The cluster analysis is<br />

<strong>the</strong>refore a means of defi n<strong>in</strong>g biotopes on <strong>the</strong> basis of fi eld<br />

measurements, ra<strong>the</strong>r than on <strong>the</strong> basis of visual survey.<br />

Figures 7.8(a) and (b) show <strong>the</strong> spatial coverage and location<br />

of <strong>the</strong> six dist<strong>in</strong>ct velocity clusters at higher fl ow stage<br />

result<strong>in</strong>g from <strong>the</strong> analysis, while Figures 7.8(c) and (d)<br />

detail <strong>the</strong> velocity characteristics associated with each<br />

cluster over a range of fl ow changes from low fl ow to a<br />

high <strong>in</strong>-bank fl ow.<br />

Several po<strong>in</strong>ts emerge from this analysis which <strong>in</strong>dicate<br />

that <strong>the</strong> method might help both understand and, <strong>in</strong> time,<br />

reduce uncerta<strong>in</strong>ties <strong>in</strong> habitat recognition and appraisal.<br />

The method is clearly sensitive to bedform amplitude and<br />

planform complexity. Even where <strong>the</strong> basic biotope structure<br />

of rivers is similar, <strong>the</strong> differ<strong>in</strong>g patterns of coverage<br />

are suffi cient to <strong>in</strong>dicate a degree of ‘biotope subtlety’ –<br />

contrast Figures 7.8(a), where fl ow organisation at high<br />

stage is marked by l<strong>in</strong>earity, and 7.8(b), where patch<strong>in</strong>ess<br />

is dom<strong>in</strong>ant. This is supported by <strong>the</strong> stage-dependent


North<strong>in</strong>g (m)<br />

Velocity (m/s)<br />

1060<br />

1040<br />

1020<br />

1000<br />

980<br />

960<br />

940<br />

920<br />

Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 119<br />

Cluster1<br />

Cluster2<br />

Cluster3<br />

Cluster4<br />

Cluster5<br />

Cluster6<br />

North<strong>in</strong>g (m)<br />

1070<br />

1060<br />

1050<br />

1040<br />

1030<br />

1020<br />

Cluster1<br />

Cluster2<br />

Cluster3<br />

Cluster4<br />

Cluster5<br />

Cluster6<br />

900<br />

4990 5000 5010 5020<br />

East<strong>in</strong>g (m)<br />

5030 5040<br />

1010<br />

4985 4990 4995<br />

East<strong>in</strong>g (m)<br />

5000 5005<br />

(a) (b)<br />

1<br />

0.8<br />

0.9<br />

0.8<br />

0.7<br />

0.7<br />

0.6<br />

0.6<br />

0.5<br />

0.5<br />

0.4<br />

0.4<br />

0.3<br />

0.3<br />

0.2<br />

0.2<br />

0.1<br />

0.1<br />

0<br />

Low ‘x’ V mean Med ‘x’ V mean High ‘x’ V mean<br />

0<br />

Low ‘x’ V mean Med ‘x’ V mean High ‘x’ V mean<br />

Stage class Stage class<br />

(c) (d)<br />

Velocity (m/s)<br />

Figure 7.8 Results of cluster analysis to determ<strong>in</strong>e coverage and behaviour of velocity ‘patches’ <strong>in</strong> <strong>the</strong> <strong>River</strong>s Cole, Birm<strong>in</strong>gham (a<br />

and c) and Tern, Shropshire (b and d) (Source: Modifi ed from Emery, 2003)<br />

velocity behaviour seen <strong>in</strong> Figures 7.8(c) and (d). In <strong>the</strong><br />

<strong>River</strong> Cole, <strong>the</strong> straighter, simpler channel, patch coherence<br />

<strong>in</strong>creases and cluster number decreases (coverage<br />

<strong>in</strong>creases) as fl ow stage rises. Thus, <strong>the</strong> river ‘simplifi es’<br />

with <strong>in</strong>creas<strong>in</strong>g fl ow stage <strong>in</strong>to three basic group<strong>in</strong>gs:<br />

channel marg<strong>in</strong>s, channel centrel<strong>in</strong>e, and areas above<br />

bedform crests. However, <strong>in</strong> <strong>the</strong> <strong>River</strong> Tern, <strong>the</strong> lower<br />

stage complexity is largely ma<strong>in</strong>ta<strong>in</strong>ed as fl ow stage rises,<br />

possibly because <strong>the</strong> fl ow now <strong>in</strong>teracts with large bankside<br />

tree roots creat<strong>in</strong>g new or ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g old, biotopes<br />

(for fur<strong>the</strong>r discussion, see Emery et al., 2003).<br />

The o<strong>the</strong>r key po<strong>in</strong>ts emerg<strong>in</strong>g from this analysis are<br />

equally important and draw attention to some pitfalls <strong>in</strong><br />

identifi cation and representation, which might improve<br />

survey methods <strong>in</strong> <strong>the</strong> future. First, <strong>in</strong> assess<strong>in</strong>g biotopes<br />

and functional habitats, <strong>the</strong> entire fl ow regime should be<br />

considered – what exists at low fl ow may or may not<br />

change <strong>in</strong> character as stage rises. Second, <strong>the</strong>re is an<br />

obvious role for hydraulic and hydrodynamic fl ow modell<strong>in</strong>g.<br />

This might support fi eld observations (which may be<br />

limited to s<strong>in</strong>gle fl ows) and assist <strong>in</strong> restoration design and<br />

appraisal, where <strong>the</strong> ability to predict and to visualise


120 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

chang<strong>in</strong>g aspects of fl ow and habitat performance as fl ow<br />

stage varies is crucial.<br />

7.4.3 Hydraulic Habitat Simulation and Modell<strong>in</strong>g<br />

Assessment of river fl ow management options, <strong>in</strong>clud<strong>in</strong>g<br />

river restoration and changes <strong>in</strong> fl ow regime, often <strong>in</strong>volves<br />

assess<strong>in</strong>g scenarios that fall outside <strong>the</strong> range of observed<br />

conditions. This negates <strong>in</strong>vestigation us<strong>in</strong>g direct analysis<br />

of fi eld observations and necessitates <strong>the</strong> use of predictive<br />

models. Models employ a variety of hydrological, morphological<br />

and hydraulic parameters to predict variations,<br />

which might <strong>the</strong>n be related to <strong>the</strong> abundance and distribution<br />

of aquatic organisms. The parameters vary accord<strong>in</strong>g<br />

to <strong>the</strong> spatial scale of <strong>the</strong> simulation. Habitat modell<strong>in</strong>g<br />

frequently <strong>in</strong>volves two steps: <strong>the</strong> physical analysis and<br />

evaluation of <strong>the</strong> hydraulic and morphological aspects of<br />

<strong>the</strong> problem, and <strong>the</strong> subsequent l<strong>in</strong>kage to ecological<br />

response.<br />

Uncerta<strong>in</strong>ties <strong>in</strong> <strong>the</strong> Hydraulic and Hydrodynamic<br />

Modell<strong>in</strong>g Process: <strong>the</strong> Fundamentals<br />

All hydraulic and hydrodynamic fl ow simulation models<br />

beg<strong>in</strong> with equations for <strong>the</strong> conservation of mass and<br />

momentum, which are modifi ed for <strong>the</strong> bed and surface<br />

boundary conditions found <strong>in</strong> shallow, open channel fl ow.<br />

As a result of limitations imposed by <strong>the</strong> scale of <strong>the</strong><br />

smaller turbulent motions, <strong>the</strong> govern<strong>in</strong>g equations are not<br />

tractable <strong>in</strong> <strong>the</strong>ir orig<strong>in</strong>al form and require additional<br />

modifi cations or ‘closure models’ to provide computational<br />

outputs. These assumptions <strong>in</strong> turn give rise to new<br />

terms <strong>in</strong> <strong>the</strong> govern<strong>in</strong>g equations, which are <strong>the</strong>mselves<br />

subsequently manipulated to satisfy <strong>the</strong> requirements of<br />

<strong>the</strong> solution, which may be represented as a one-, two- or<br />

three-dimensional numerical scheme (for full review, see<br />

Lane, 1998). Generally, <strong>in</strong> eco-hydraulic and river restoration<br />

applications, <strong>the</strong> uncerta<strong>in</strong>ties <strong>in</strong> model derivation are<br />

secondary to those of model application to <strong>the</strong> particular<br />

case study. In effect, modell<strong>in</strong>g is really a recursive<br />

process, <strong>in</strong>volv<strong>in</strong>g uncerta<strong>in</strong>ties at all stages from premodel<br />

data collection, through model verifi cation dur<strong>in</strong>g<br />

<strong>the</strong> numerical calculation, to post-modell<strong>in</strong>g validation,<br />

where results are compared with fi eld measurements and<br />

o<strong>the</strong>r sources of evidence/expertise to assess, or appraise<br />

<strong>the</strong> model performance.<br />

Traditionally, hydraulic modell<strong>in</strong>g has been focused on<br />

one-dimensional representation of open channel fl ow<br />

(Chow, 1973), <strong>in</strong> which <strong>the</strong> cross-sectionally averaged<br />

depth and velocity are obta<strong>in</strong>ed. These models may be<br />

used to simulate <strong>the</strong> passage of fl oods through reaches and<br />

channel systems, or for <strong>the</strong> assessment of channel stability<br />

given particular boundary characteristics (see Chapter 5<br />

for more details). In ecological applications, <strong>the</strong>y may be<br />

of use to determ<strong>in</strong>e depth and tim<strong>in</strong>g of <strong>in</strong>undation associated<br />

with particular events. Hydrographs are routed<br />

through cross-sections ei<strong>the</strong>r by runn<strong>in</strong>g a series of steadystate<br />

solutions for different discharge ‘b<strong>in</strong>s’ <strong>in</strong> <strong>the</strong> hydrograph<br />

or us<strong>in</strong>g unsteady methods to model <strong>the</strong> hydrograph.<br />

Common one-dimensional implementations of this<br />

sort are iSIS, HEC–RAS and MIKE11, while <strong>the</strong> most<br />

popular hydraulically-coupled habitat suitability model –<br />

PHABSIM (Spence and Hickley, 2000, and below) – uses<br />

one-dimenstional approaches to predict velocity and depth<br />

<strong>in</strong> channel cells or slices, defi ned between adjacent measured<br />

cross-sections. In <strong>the</strong>ir simplest forms, <strong>the</strong>se models<br />

rely on <strong>the</strong> ability to specify <strong>the</strong> relationship between<br />

depth, water surface slope and velocity via an empirical<br />

‘constant’ known as Mann<strong>in</strong>g’s n, which is normally<br />

assumed to represent <strong>the</strong> fl ow resistance aris<strong>in</strong>g from<br />

boundary friction of <strong>the</strong> channel. The value of n is usually<br />

determ<strong>in</strong>ed from look-up tables (e.g. Chow, 1973). Coeffi<br />

cients of channel expansion and contraction may be used<br />

<strong>in</strong> addition to channel geometry to account for additional<br />

components of energy loss aris<strong>in</strong>g from fl ow acceleration<br />

and deceleration. Away from channels of very simple<br />

geometry and boundary characteristics, <strong>the</strong> appropriate<br />

determ<strong>in</strong>ation of n is more subtle: to ma<strong>in</strong>ta<strong>in</strong> predictive<br />

success, for example, n is varied <strong>in</strong>versely with depth and,<br />

<strong>in</strong> many situations, n is really a compound calibration<br />

factor express<strong>in</strong>g (with more-or-less success) <strong>the</strong> various<br />

contributions to fl ow resistance, and hence energy loss <strong>in</strong><br />

channels: gra<strong>in</strong>, bedform, and spill (channel curvature).<br />

Some of <strong>the</strong> issues <strong>in</strong> more complex modell<strong>in</strong>g are outl<strong>in</strong>ed<br />

below.<br />

Two- and three-dimensional Numerical Schemes <strong>in</strong><br />

Ecological <strong>Restoration</strong> Applications<br />

Ideally, two- and three-dimensional approaches are required<br />

for realistic simulation of fl ow behaviour and to better represent<br />

<strong>the</strong> diversity and variability of physical habitats<br />

(Crowder and Diplas, 2000), thus overcom<strong>in</strong>g <strong>the</strong> limitations<br />

of popular habitat simulation models. In particular,<br />

account<strong>in</strong>g for small scale stage-dependent variations <strong>in</strong><br />

fl ow is important <strong>in</strong> <strong>the</strong> ma<strong>in</strong>tenance and survival of <strong>the</strong><br />

biotic community. In this way, more complex numerical<br />

fl ow simulation offers potential as a ‘support<strong>in</strong>g’ function<br />

for river rehabilitation schemes. Two- and threedimensional<br />

codes are now widely accessible and can be<br />

run over a wide range of discharge conditions, <strong>in</strong>clud<strong>in</strong>g<br />

over-bank fl ows, but <strong>the</strong>re are numerous sources of uncerta<strong>in</strong>ty<br />

at all stages of <strong>the</strong> modell<strong>in</strong>g process associated with<br />

<strong>the</strong>se simulations. Models <strong>in</strong>volve many underly<strong>in</strong>g


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 121<br />

assumptions at <strong>the</strong>ir <strong>in</strong>put stage, <strong>the</strong>y are sensitive to <strong>the</strong><br />

numerical structures <strong>in</strong>volved <strong>in</strong> obta<strong>in</strong><strong>in</strong>g <strong>the</strong> fl ow solution<br />

and outputs may be diffi cult to observe or measure <strong>in</strong><br />

<strong>the</strong> fi eld. Some of <strong>the</strong> more common models, model applications<br />

and model uncerta<strong>in</strong>ties are listed <strong>in</strong> Table 7.3.<br />

Perhaps <strong>the</strong> most basic but much neglected area of<br />

uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> application of more complex models<br />

rests with <strong>the</strong> fi eld data requirements for model calibration<br />

and validation. This is most onerous where modell<strong>in</strong>g is<br />

used at a spatial resolution commensurate with sub-reach<br />

detail <strong>in</strong> habitat <strong>in</strong> conditions of complex topography<br />

(Crowder and Diplas, 2000). These have led some to argue<br />

that <strong>the</strong> ecological and hydraulic/hydrodynamic approaches<br />

should proceed <strong>in</strong>dependently, because river environments<br />

are ‘too complex’ for a coupled approach (Kondolf et al.,<br />

2000). Field observations <strong>in</strong>dicate that physical habitat<br />

may be structured at an extremely small scale, such as<br />

around boulders, or extremely close to channel banks<br />

(Railsback et al., 1999), which necessitate huge <strong>in</strong>creases<br />

<strong>in</strong> model effort and computational time, and whose characteristics<br />

may not, <strong>in</strong> any case, be picked up by conventional<br />

fi eld survey. Figure 7.9 for example, illustrates how<br />

our knowledge of <strong>the</strong> character of <strong>the</strong> river changes as <strong>the</strong><br />

number and density of fi eld measurements at various fl ow<br />

stages is <strong>in</strong>creased or decreased.<br />

Here, <strong>the</strong> results of an orig<strong>in</strong>al fi eld survey of more than<br />

350 data po<strong>in</strong>ts obta<strong>in</strong>ed by survey through a 150 m river<br />

reach of <strong>the</strong> <strong>River</strong> Cole (represent<strong>in</strong>g both longitud<strong>in</strong>al<br />

and cross section variation with an average po<strong>in</strong>t-to-po<strong>in</strong>t<br />

spac<strong>in</strong>g of 1–2 m) have been plotted as a frequency histogram<br />

of velocities, and <strong>the</strong>n subsequently re-plotted <strong>in</strong> <strong>the</strong><br />

same way but with <strong>the</strong> orig<strong>in</strong>al dataset systematically<br />

degraded by a factor of one half or one third. The resultant<br />

changes thus give an idea of <strong>the</strong> effects of vary<strong>in</strong>g <strong>the</strong> fi eld<br />

data collection efforts or schemes <strong>in</strong> characteris<strong>in</strong>g ‘true’<br />

fi eld velocities, both prior to modell<strong>in</strong>g simulation and <strong>in</strong><br />

post-modell<strong>in</strong>g assessment. Several po<strong>in</strong>ts are worthy of<br />

note from this analysis. Firstly, channel velocity characteristics<br />

differ depend<strong>in</strong>g upon fl ow stage: at lower fl ow<br />

stage, results are more skewed than at higher stage, but at<br />

<strong>the</strong> higher stage <strong>the</strong>re is <strong>the</strong> suggestion of multiple modes<br />

<strong>in</strong> an o<strong>the</strong>rwise more even velocity distribution. This<br />

change <strong>in</strong> velocity distribution should be <strong>in</strong>terpreted <strong>in</strong><br />

conjunction with <strong>the</strong> results of <strong>the</strong> biotope analysis <strong>in</strong><br />

Section 7.4.2, s<strong>in</strong>ce <strong>the</strong> fi eld data are common to both. In<br />

this case, it seems that, as biotopes ‘clarify’, <strong>the</strong> distribution<br />

of velocities also becomes organised about dom<strong>in</strong>ant<br />

peaks, refl ect<strong>in</strong>g <strong>the</strong> <strong>in</strong>fl uence of channel marg<strong>in</strong>s, channel<br />

centrel<strong>in</strong>es and shallows over bedform crests. The second<br />

po<strong>in</strong>t to note is that sub-sampl<strong>in</strong>g from <strong>the</strong> orig<strong>in</strong>al data<br />

Table 7.3 Common numerical fl ow models, <strong>the</strong>ir pr<strong>in</strong>cipal applications <strong>in</strong> river restoration and eco-hydraulics, and some issues<br />

relat<strong>in</strong>g to parameterisation and model <strong>in</strong>terpretation<br />

Class of model<br />

and common<br />

implementation<br />

One-dimensionnal<br />

HEC-RAS<br />

ISIS<br />

backwater<br />

Two-dimensionnal<br />

Telemac<br />

RMA<br />

Three-dimensionnal<br />

SSIIM<br />

Fluent<br />

CFX<br />

Pr<strong>in</strong>cipal application <strong>in</strong><br />

channel restoration<br />

fl ood rout<strong>in</strong>g and channel<br />

conveyance<br />

<strong>in</strong>itial water surface<br />

elevation (<strong>in</strong> higher order<br />

models); velocity and<br />

depth between cross<br />

sections <strong>in</strong> PHABSIM<br />

dynamic simulation of<br />

channel fl ow and<br />

fl oodpla<strong>in</strong> <strong>in</strong>undation;<br />

cross-sectional patterns<br />

of habitat suitability<br />

detailed sub-reach<br />

modell<strong>in</strong>g of habitat<br />

suitability<br />

Required parameters Pr<strong>in</strong>cipal aspects of uncerta<strong>in</strong>ty<br />

downstream discharge;<br />

upstream water level;<br />

Mann<strong>in</strong>g’s n; channel<br />

expansion and contraction<br />

coeffi cients; basic channel<br />

cross section morphology<br />

boundary roughness (ks);<br />

fl ow <strong>in</strong>formation as above;<br />

bed topography; turbulence<br />

closure model; choice of<br />

numerical solver and<br />

relaxation coeffi cients;<br />

correct estimation of n (particularly with<br />

depth and vegetation changes) and<br />

channel expansion and contraction<br />

coeffi cients; loss of <strong>in</strong>formation on<br />

cross-sectional fl ow distribution; no<br />

account of channel curvature<br />

loss of <strong>in</strong>formation on depth-related fl ow<br />

properties; time-stepp<strong>in</strong>g <strong>in</strong> unsteady<br />

solutions; poor representation of<br />

secondary fl ow from channel<br />

discont<strong>in</strong>uities; mesh<strong>in</strong>g issues;<br />

representation of wett<strong>in</strong>g and dry<strong>in</strong>g<br />

as above most of <strong>the</strong> above, plus: meshdependence<br />

<strong>in</strong> <strong>the</strong> vertical as well as<br />

cross-section; determ<strong>in</strong>ation of <strong>the</strong> free<br />

water surface


122 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Frequency (%)<br />

Frequency (%)<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

Resultant velocity (m 3 s –1 )<br />

(a)<br />

full data set<br />

half data set<br />

third data set<br />

0<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4<br />

Resultant velocity (m 3 s –1 )<br />

(b)<br />

full data set<br />

half data set<br />

third data set<br />

Figure 7.9 Effect of degrad<strong>in</strong>g <strong>the</strong> velocity sample on <strong>the</strong> distribution of depth average velocities: a) low fl ow; b) high <strong>in</strong>-bank<br />

fl ow


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 123<br />

does not have an even effect <strong>in</strong> all velocity classes. This<br />

may or may not be of signifi cance but is likely to impart<br />

uncerta<strong>in</strong>ty <strong>in</strong>to <strong>the</strong> assessment of model performance if<br />

<strong>the</strong> affected velocity classes are those judged to be of most<br />

relevance for habitat suitability. This arises because chang<strong>in</strong>g<br />

<strong>the</strong> frequency of particular velocity class observations<br />

is equivalent to under- or over-represent<strong>in</strong>g <strong>the</strong> spatial<br />

coverage associated with that velocity. For fur<strong>the</strong>r discussion,<br />

see Clifford et al. (2002a, 2002b).<br />

Once acceptable fi eld velocity data have been obta<strong>in</strong>ed,<br />

<strong>the</strong>se must be coupled with representations of boundary<br />

topography and channel geometry. This should be suffi -<br />

ciently detailed to <strong>in</strong>corporate both small-scale irregularities<br />

<strong>in</strong> bed and banks, <strong>the</strong> major breaks of slope associated<br />

with channel marg<strong>in</strong>s and more subtle fl oodpla<strong>in</strong> topography.<br />

Breakl<strong>in</strong>es are notoriously diffi cult to handle, and if<br />

fi ne-scale survey is unavailable <strong>the</strong>n topography has to be<br />

<strong>in</strong>terpolated, <strong>in</strong>troduc<strong>in</strong>g a fur<strong>the</strong>r source of uncerta<strong>in</strong>ty,<br />

which also occurs <strong>in</strong> areas of low relief, too. Simple l<strong>in</strong>ear<br />

<strong>in</strong>terpolation, for example, may be preferable to avoid<br />

spurious high- and low-po<strong>in</strong>ts (French and Clifford, 2000).<br />

Topography forms <strong>the</strong> template of <strong>the</strong> modell<strong>in</strong>g grid or<br />

mesh which is draped over <strong>the</strong> boundary surface and which<br />

forms <strong>the</strong> basis for <strong>the</strong> numerical solution to <strong>the</strong> govern<strong>in</strong>g<br />

fl ow equations. The form of <strong>the</strong> mesh depends upon <strong>the</strong><br />

k<strong>in</strong>d of model scheme adopted. Meshes might exploit triangular<br />

(<strong>in</strong> fi nite element schemes) or rectangular elements<br />

(<strong>in</strong> fi nite difference and fi nite volume schemes), of<br />

ei<strong>the</strong>r constant or variable size. Triangular meshes are<br />

more versatile, but most ecological applications use rectangular<br />

grid elements and structured grid schemes. While<br />

allow<strong>in</strong>g elements to vary <strong>in</strong> size, <strong>the</strong>se always require <strong>the</strong><br />

same number of elements <strong>in</strong> each cross plane of solution.<br />

In practice, <strong>the</strong> constra<strong>in</strong>ts of gridd<strong>in</strong>g normally require<br />

some degradation <strong>in</strong> <strong>the</strong> quality of <strong>the</strong> topographic representation,<br />

s<strong>in</strong>ce multiply<strong>in</strong>g <strong>the</strong> number of very small elements<br />

to represent more detailed topographic variation<br />

imparts a disproportionately large <strong>in</strong>crement <strong>in</strong> computer<br />

time required to solve <strong>the</strong> problem and may result <strong>in</strong><br />

<strong>in</strong>stabilities <strong>in</strong> <strong>the</strong> solution. Many of <strong>the</strong>se issues receive<br />

a thorough exploration <strong>in</strong> Anderson et al. (1996). As computer<br />

power has <strong>in</strong>creased, more detailed modell<strong>in</strong>g is<br />

possible, but approach<strong>in</strong>g closer representations of reality<br />

does not simply depend upon enhanced computer speed<br />

or memory. A key issue is that, with smaller elements, bed<br />

gra<strong>in</strong>s and micro-topography may approach or exceed <strong>the</strong><br />

size of model elements, render<strong>in</strong>g solution impossible.<br />

This is especially pert<strong>in</strong>ent <strong>in</strong> two-dimensional modell<strong>in</strong>g,<br />

where <strong>the</strong> number and size of elements <strong>in</strong> <strong>the</strong> vertical may<br />

also be varied, and <strong>the</strong> issue is related to <strong>the</strong> wider question<br />

of <strong>the</strong> appropriate parameterisation of multiple scales of<br />

fl ow resistance (Nicholas, 2001).<br />

With respect to fl ow resistance, most two-dimensional<br />

models often rema<strong>in</strong> reliant on a boundary resistance<br />

value specifi ed by n or by Chezy’s conveyance coeffi cient,<br />

C. Three-dimensional models employ a boundary roughness<br />

specifi ed <strong>in</strong> terms of an equivalent roughness length,<br />

k s, which propagates through <strong>the</strong> boundary cells. While ks<br />

can be determ<strong>in</strong>ed <strong>in</strong> relation to <strong>the</strong> boundary gra<strong>in</strong> size,<br />

<strong>the</strong> numerical value required to obta<strong>in</strong> plausible solutions<br />

is generally much larger than gra<strong>in</strong> size alone would imply,<br />

giv<strong>in</strong>g rise to uncerta<strong>in</strong>ty <strong>in</strong> its <strong>in</strong>terpretation. Flow resistance<br />

effects beyond sk<strong>in</strong> friction may be <strong>in</strong>corporated<br />

witt<strong>in</strong>gly or unwitt<strong>in</strong>gly <strong>in</strong>to <strong>the</strong> mesh<strong>in</strong>g strategy – for<br />

example, by vary<strong>in</strong>g <strong>the</strong> size of cells around <strong>the</strong> boundary<br />

as compared to <strong>the</strong> ma<strong>in</strong> body of <strong>the</strong> fl ow (Clifford et al.,<br />

2002a). O<strong>the</strong>r areas of uncerta<strong>in</strong>ty <strong>in</strong> both model application<br />

and result <strong>in</strong>terpretation <strong>in</strong>clude: whe<strong>the</strong>r <strong>the</strong> model<br />

uses a fi xed water surface (lid) or allows <strong>the</strong> water<br />

surface to freely adjust; <strong>the</strong> scheme used to allow <strong>the</strong><br />

simulation of turbulence effects (<strong>the</strong> closure assumptions);<br />

and <strong>the</strong> complexity of <strong>the</strong> solver (which determ<strong>in</strong>es how<br />

many and which neighbour<strong>in</strong>g po<strong>in</strong>ts are considered <strong>in</strong> <strong>the</strong><br />

propagation of <strong>the</strong> numerical solution through <strong>the</strong> mesh).<br />

Many of <strong>the</strong>se issues are common to both two- and threedimensional<br />

schemes, although similar strategies and<br />

choices will not necessarily have <strong>the</strong> same effects <strong>in</strong> <strong>the</strong><br />

differ<strong>in</strong>g model implementations. Nicholas (2001) provides<br />

a comprehensive start<strong>in</strong>g po<strong>in</strong>t for assess<strong>in</strong>g and<br />

manag<strong>in</strong>g <strong>the</strong>se uncerta<strong>in</strong>ties. The particular limitation of<br />

two-dimentional models is <strong>the</strong> parameterisation of secondary<br />

circulation effects on momentum transport, which<br />

has concentrated on <strong>the</strong> effects of channel curvature, but<br />

not on topographic discont<strong>in</strong>uities and river confl uences<br />

(Lane, 1998).<br />

Rarely do <strong>the</strong> uncerta<strong>in</strong>ties aris<strong>in</strong>g with respect to each<br />

aspect of <strong>the</strong> modell<strong>in</strong>g occur <strong>in</strong>dependently: Figure 7.10<br />

for example, illustrates <strong>the</strong> chang<strong>in</strong>g fl ow representations<br />

aris<strong>in</strong>g from coarser- and fi ner-scale element sizes us<strong>in</strong>g<br />

<strong>the</strong> SSIMM model of <strong>the</strong> <strong>River</strong> Cole, near Birm<strong>in</strong>gham,<br />

UK. Apart from <strong>the</strong> element size, all o<strong>the</strong>r model parameters<br />

are held constant. As <strong>the</strong> results demonstrate, modelled<br />

velocities vary <strong>in</strong> both planform and <strong>in</strong> cross-section<br />

and vertical distributions as grid element size is varied<br />

from approximately 2–0.5 m. The coarse grid distorts<br />

<strong>the</strong> channel shape from generally rectangular to a trapezoidal<br />

cross section and this contributes to bias<strong>in</strong>g <strong>the</strong> faster<br />

fl ow towards <strong>the</strong> channel centre, rais<strong>in</strong>g velocities above<br />

those measured <strong>in</strong> <strong>the</strong> fi eld. This effect is also enhanced<br />

by <strong>the</strong> <strong>in</strong>teraction of grid element size and boundary<br />

roughness, which <strong>in</strong> <strong>the</strong> model extends to 20% of <strong>the</strong> near<br />

boundary grid element. When elements are large, a larger<br />

proportion of <strong>the</strong> channel is thus affected by <strong>the</strong> boundary<br />

roughness, even though <strong>the</strong> numerical value of k s is <strong>the</strong>


124 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

HIGH flow<br />

coarse grid<br />

Pool<br />

Pool<br />

1.4<br />

1.3<br />

1.2<br />

1.1<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

1.4<br />

1.3<br />

1.2<br />

1.1<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Modelled<br />

velocities<br />

Modelled<br />

velocities<br />

Pool<br />

HIGH flow<br />

f<strong>in</strong>e grid<br />

Pool<br />

Riffle crest<br />

Riffle crest<br />

(m)<br />

(m)<br />

1060<br />

1040<br />

1020<br />

1000<br />

980<br />

960<br />

940<br />

1060<br />

1040<br />

1020<br />

1000<br />

980<br />

960<br />

940<br />

N<br />

920<br />

4980 5000<br />

(m)<br />

5020 5040<br />

N<br />

PT 4<br />

Flow<br />

PT 4<br />

Flow<br />

PT 2<br />

PT 2<br />

RC 4<br />

PT 3<br />

RC 4<br />

RC 2<br />

PT 3<br />

RC 2<br />

RC 3<br />

RC 3<br />

PT 1<br />

RC 1<br />

PT 1<br />

RC 1<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

-0.05<br />

-0.1<br />

-0.15<br />

-0.2<br />

-0.25<br />

-0.3<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

-0.05<br />

-0.1<br />

-0.15<br />

-0.2<br />

-0.25<br />

-0.3<br />

920<br />

4980 5000<br />

(m)<br />

5020 5040<br />

Modelledfield<br />

velocities<br />

Modelledfield<br />

velocities<br />

Figure 7.10 (See also colour plate section) Chang<strong>in</strong>g SSIIM model output and model ‘fi t’ with as a function of mesh element size<br />

<strong>in</strong> a reach of <strong>the</strong> <strong>River</strong> Cole, near Birm<strong>in</strong>gham, UK


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 125<br />

same. This effective <strong>in</strong>crease <strong>in</strong> friction comb<strong>in</strong>ed with <strong>the</strong><br />

reduced area of <strong>the</strong> channel <strong>in</strong> <strong>the</strong> coarser grid scheme<br />

‘push’ <strong>the</strong> fl ow to <strong>the</strong> centrel<strong>in</strong>e. In <strong>the</strong> streamwise direction,<br />

<strong>the</strong> coarser grid resolution is also <strong>in</strong>suffi cient to represent<br />

<strong>the</strong> ‘patch<strong>in</strong>ess’ of velocity seen <strong>in</strong> <strong>the</strong> fi ner grid<br />

case and, hence, both misses and distorts <strong>the</strong> spatial distribution<br />

of potential physical habitat. By contrast, use of<br />

<strong>the</strong> fi ner mesh over-emphasises velocity patch<strong>in</strong>ess and<br />

gives rise to problems of near-bank velocity representation.<br />

Here, apparently high velocity values cannot be corroborated<br />

with fi eld data which were collected at coarser<br />

resolution.<br />

Some models allow adaptive grids which vary <strong>in</strong> complexity<br />

as <strong>the</strong> problem evolves. The most fundamental<br />

illustration of this is <strong>the</strong> ability (or o<strong>the</strong>rwise) to cope with<br />

wett<strong>in</strong>g and dry<strong>in</strong>g, which will occur as fl ow stage rises<br />

and <strong>in</strong>undates fl oodpla<strong>in</strong> or channel marg<strong>in</strong>s, before reced<strong>in</strong>g.<br />

Full simulation of this requires an unsteady solution<br />

coupled to one of several schemes to ei<strong>the</strong>r <strong>in</strong>clude or<br />

exclude fully or partially dry elements. Dynamic solutions<br />

are generally confi ned to two-dimensional schemes. A<br />

simpler solution is to produce a number of grids, each of<br />

which is fully wet for <strong>the</strong> particular fl ow stage under consideration.<br />

In analogy with one-dimensional models, <strong>the</strong><br />

hydrograph may <strong>the</strong>n be simulated from a series of steady<br />

state solutions.<br />

Much of <strong>the</strong> uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> modell<strong>in</strong>g process may<br />

be summarised by <strong>the</strong> terms verifi cation and validation.<br />

These are fundamental considerations <strong>in</strong> numerical fl ow<br />

modell<strong>in</strong>g but have recently been re-exam<strong>in</strong>ed as <strong>the</strong><br />

number of models and range of model application has<br />

burgeoned. Verifi cation relates to <strong>the</strong> ability to correctly<br />

solve <strong>the</strong> appropriate equations of <strong>the</strong> model problem at<br />

hand (i.e. whe<strong>the</strong>r <strong>the</strong> solution is an accurate solution of<br />

<strong>the</strong> particular choice of model); whereas validation relates<br />

to <strong>the</strong> plausibility of <strong>the</strong> model as a whole and <strong>the</strong> test<strong>in</strong>g<br />

of parameters predicted by <strong>the</strong> model (thus <strong>in</strong>volv<strong>in</strong>g<br />

<strong>the</strong> degree of fi t between prediction and measurement).<br />

As model applications have <strong>in</strong>creased, <strong>the</strong> dist<strong>in</strong>ctions<br />

between <strong>the</strong> two have become more blurred, and <strong>the</strong> term<br />

model assessment or appraisal has been used (Lane and<br />

Richards, 2001). Hardy et al. (2003) follow conventional<br />

eng<strong>in</strong>eer<strong>in</strong>g practice, argu<strong>in</strong>g that verifi cation is <strong>the</strong> essential<br />

element <strong>in</strong> <strong>the</strong> assessment process, which must precede<br />

attempts at validation. However, relatively little attention<br />

has been given to develop<strong>in</strong>g means of model assessment<br />

that are tailored to <strong>the</strong> application, particularly with respect<br />

to eco-hydraulics. In environmental systems, for example,<br />

<strong>the</strong> degree of experimental closure is much less than <strong>in</strong><br />

laboratory conditions or eng<strong>in</strong>eer<strong>in</strong>g problems, and <strong>the</strong><br />

application of models is as much (if not more) to produce<br />

data to support or <strong>in</strong>form new fi eld strategies or explana-<br />

tory <strong>in</strong>terpretations/<strong>in</strong>sights (Oreskes et al., 1994) as is it<br />

is to judge <strong>the</strong> correspondence between measured and<br />

modelled values. Thus, <strong>in</strong> ecological applications, where<br />

<strong>the</strong> uncerta<strong>in</strong>ties of measurements (both physical and biological)<br />

are coupled with an <strong>in</strong>tr<strong>in</strong>sic dynamism and variability,<br />

Clifford et al. (2005) have argued that assessment<br />

of models necessarily requires a more ‘relaxed’ approach.<br />

This is illustrated <strong>in</strong> Figure 7.11, where planform representations<br />

of model:fi eld correspondence of velocities for<br />

<strong>the</strong> <strong>River</strong> Cole are shown.<br />

In Figure 7.11(a), <strong>the</strong> fi t of <strong>the</strong> model is shown for a<br />

higher fl ow stage, purely as <strong>the</strong> absolute value of <strong>the</strong> po<strong>in</strong>tby-po<strong>in</strong>t<br />

modelled-fi eld velocity. On this basis, approximately<br />

80% of <strong>the</strong> channel area is modelled to with<strong>in</strong><br />

±0.2 ms −1 , and approximately 40% to with<strong>in</strong> ±0.1 ms −1<br />

(Figure 7.11(c)). Figure 7.11(b), however, shows <strong>the</strong><br />

improvement <strong>in</strong> fi t when a degree of ‘relaxation’ is applied.<br />

In this case, velocity ‘errors’ were recoded to zero<br />

provided that two criteria were met: fi rst, <strong>the</strong> modelled<br />

velocity must lie with<strong>in</strong> 0.1 ms −1 of <strong>the</strong> fi eld velocity and,<br />

second, this 0.1 ms −1 difference must also lie with<strong>in</strong> a 1 m<br />

radius of <strong>the</strong> actual fi eld measurement. By allow<strong>in</strong>g this<br />

relaxation, <strong>the</strong> fi t of <strong>the</strong> model is dramatically improved:<br />

reference to Figure 7.11(c) shows that only approximately<br />

5% of <strong>the</strong> channel fails to fulfi l <strong>the</strong>se jo<strong>in</strong>t criteria. The<br />

justifi cation for <strong>the</strong> relaxation lies <strong>in</strong> <strong>the</strong> fact that: (a) fi eld<br />

measurements are rarely precise enough to warrant absolute<br />

comparison to modelled results, particularly where<br />

both fi eld and modelled velocities have been <strong>in</strong>terpolated<br />

to <strong>the</strong> same grid for comparison; (b) <strong>in</strong> ecological applications,<br />

<strong>the</strong> uncerta<strong>in</strong>ties <strong>in</strong> relat<strong>in</strong>g physical parameters to<br />

ecological response are large; and (c) <strong>in</strong> any case plant and<br />

animal communities are mobile, dynamic communities,<br />

ra<strong>the</strong>r than static and adapt to exploit <strong>the</strong> most favourable<br />

environments. Ecological simulation conta<strong>in</strong>s, <strong>the</strong>n, an<br />

essential uncerta<strong>in</strong>ty, which might usefully be <strong>in</strong>corporated<br />

<strong>in</strong>to numerical model assessment so as to improve<br />

<strong>in</strong>terpretation!<br />

Ecological Uncerta<strong>in</strong>ties <strong>in</strong> Eco-hydraulic Modell<strong>in</strong>g<br />

Ecological uncerta<strong>in</strong>ties relate ma<strong>in</strong>ly to <strong>the</strong> simplifi cation,<br />

observation and test<strong>in</strong>g of habitat suitability criteria.<br />

Most frequently, <strong>the</strong>se have been characterised <strong>in</strong> terms of<br />

species’ preferences or abundances, result<strong>in</strong>g <strong>in</strong> Habitat<br />

Suitability Indices (HSIs), although more recent approaches<br />

seek l<strong>in</strong>kages to particular aspects of life cycle and growth<br />

of organisms or communities (see <strong>the</strong> later section on<br />

bioenergetics <strong>in</strong> this chapter). Preference criteria may be<br />

simple (univariate) or complex (multivariate) and may be<br />

related to <strong>the</strong> physically-determ<strong>in</strong>ed environment through<br />

correspondence (association) or through <strong>the</strong> application of


126 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

(m)<br />

Cumulative Area (%)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

-1<br />

-0.8<br />

(a)<br />

PT1<br />

RC 1<br />

1.2<br />

1.1<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

20 5040<br />

relaxed model<br />

orig<strong>in</strong>al model<br />

-0.6<br />

-0.4<br />

-0.2<br />

(m)<br />

1060<br />

1040<br />

1020<br />

1000<br />

980<br />

960<br />

940<br />

N<br />

PT4<br />

Flow<br />

PT 2<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

-0.2<br />

-0.3<br />

920<br />

4980 5000 5020 5040<br />

(m)<br />

Modelled - Field Resultant Velocities (m s –1 )<br />

(c)<br />

orig<strong>in</strong>al model<br />

relaxed model<br />

0<br />

0.2<br />

0.4<br />

0.6<br />

(b)<br />

RC 4<br />

PT 3<br />

Model<br />

RC 2<br />

RC 3<br />

PT1<br />

Relaxed model<br />

Figure 7.11 Spatial (a and b) and areal (c) representations of model fi t to fi eld measurements for <strong>the</strong> <strong>River</strong> Cole, Birm<strong>in</strong>gham<br />

(Source: Based upon Figures 5 and 6 of Clifford et al., 2005)<br />

0.8<br />

RC 1<br />

1


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 127<br />

‘rules’ which allow a degree of uncerta<strong>in</strong>ty as an essential<br />

element of <strong>the</strong> procedure. Once physical habitat and ecological<br />

response are obta<strong>in</strong>ed, results of <strong>the</strong> simulation are<br />

expressed <strong>in</strong> graphical or tabular form as values of <strong>the</strong><br />

output between limits or above thresholds, often with a<br />

spatial component, expressed as Weighted Useable Area<br />

(WUA). WUA is an aggregate measure of physical habitat<br />

quality and quantity and will be specifi c to a particular<br />

discharge and species/life stage. However, it is also clear<br />

that patterns or responses <strong>in</strong> ecology frequently result<br />

from persistent variability and renewal of opportunities,<br />

ra<strong>the</strong>r than <strong>the</strong> ‘static’ presence/absence of clear habitat<br />

delimiters. There have, <strong>the</strong>refore, been calls to explore <strong>the</strong><br />

capacity of eco-hydraulic modell<strong>in</strong>g to identify capacity,<br />

opportunity and variability ra<strong>the</strong>r than <strong>the</strong> prediction of<br />

exclusion via <strong>in</strong>dividual tolerance boundaries (Bergen<br />

et al., 2001, and Reynolds, 2002).<br />

In Europe, <strong>the</strong> Water Framework Directive will require<br />

river management at <strong>the</strong> catchment scale (Logan & Furse,<br />

2002) and this is be<strong>in</strong>g mirrored <strong>in</strong> national legislation<br />

and subsequent water management activity by organisations,<br />

such as <strong>the</strong> Environment Agency of England and<br />

Wales. Therefore, a major research question for habitat<br />

modell<strong>in</strong>g is whe<strong>the</strong>r exist<strong>in</strong>g reached-based methods<br />

can be scaled up, or whe<strong>the</strong>r an entirely new approach<br />

needs to be developed. Physical habitat modell<strong>in</strong>g <strong>in</strong>vestigations<br />

are typically confi ned to short lengths of river,<br />

approximately 50–200 m. A representative reach approach<br />

may be taken <strong>in</strong> which a reach is subjectively chosen<br />

to represent a longer length of river, <strong>in</strong>clud<strong>in</strong>g <strong>the</strong><br />

direct proportions of habitats with<strong>in</strong> that reach. A habitat<br />

mapp<strong>in</strong>g approach entails classify<strong>in</strong>g and record<strong>in</strong>g habitat<br />

types (pools, riffl es, glides etc) over long lengths of<br />

river and <strong>the</strong>n choos<strong>in</strong>g cross-sections to represent <strong>the</strong><br />

identifi ed habitat types. The results from each crosssection<br />

are <strong>the</strong>n weighted accord<strong>in</strong>g to <strong>the</strong> proportions<br />

of <strong>the</strong> identifi ed habitat types (Morhardt et al., 1983;<br />

Maddock, 1999).<br />

Progress towards catchment-scale modell<strong>in</strong>g has been<br />

made <strong>in</strong> associated fi elds such as fl ood modell<strong>in</strong>g, which<br />

provides broad-scale hydraulic predictions. This provides<br />

a potential basis for catchment-scale physical habitat<br />

assessment us<strong>in</strong>g <strong>in</strong>puts of discharge, water levels and<br />

channel geometry supplied from one-dimensional hydraulic<br />

models. Velocity variations across each cross-section<br />

can <strong>the</strong>n be predicted us<strong>in</strong>g disaggregation algorithms<br />

such as that used <strong>in</strong> <strong>the</strong> HEC–RAS model (US Army<br />

Corps of Eng<strong>in</strong>eers, 2002). Velocity and depth are used to<br />

assess habitat quality for target species. A major issue<br />

when apply<strong>in</strong>g this method is <strong>the</strong> loss of local detail when<br />

scal<strong>in</strong>g up. Booker et al. (2004a) describe <strong>the</strong> application<br />

of a method to <strong>the</strong> <strong>River</strong> Itchen, UK. This was able to<br />

predict physical habitat for an entire catchment under different<br />

water use strategy scenarios compared with present<br />

and naturalised situations.<br />

Some of <strong>the</strong> many (and complex) uncerta<strong>in</strong>ties <strong>in</strong> ecohydraulic<br />

modell<strong>in</strong>g are explored below <strong>in</strong> <strong>the</strong> context<br />

of <strong>the</strong> most commonly applied numerical scheme, <strong>the</strong><br />

PHABSIM model.<br />

7.4.4 The PHABSIM Model<br />

The fi rst step <strong>in</strong> formulat<strong>in</strong>g a habitat modell<strong>in</strong>g approach<br />

for rivers was published by Waters (1976). This led to <strong>the</strong><br />

more formal description of a computer model called<br />

PHABSIM (Physical Habitat Simulation) by <strong>the</strong> US Fish<br />

and Wildlife Service (Bovee, 1982; Bovee et al., 1998;<br />

Milhous, 1999). The PHABSIM system (Bovee, 1982;<br />

Bovee et al., 1998) is a suite of numerical models that<br />

allows quantifi cation of physical habitat for a given site,<br />

defi ned <strong>in</strong> terms of <strong>the</strong> comb<strong>in</strong>ation of depth, velocity<br />

and substrate/cover at a particular discharge (e.g. Johnson<br />

et al., 1993; Elliott et al., 1996). This system is most<br />

commonly used to assess <strong>the</strong> availability of suitable habitat<br />

for fi sh, although macro<strong>in</strong>vertebrates (Gore et al., 1998)<br />

and macrophytes (Hearne et al., 1994), which have measurable<br />

physical habitat requirements, have also been<br />

<strong>the</strong> focus of PHABSIM studies. In <strong>the</strong> United K<strong>in</strong>gdom<br />

<strong>the</strong> method has been used to assess changes <strong>in</strong> physical<br />

habitat associated with alterations <strong>in</strong> fl ow regime; for<br />

example, application to <strong>the</strong> <strong>River</strong>s Allen (Johnson et al.,<br />

1995), Piddle (Strevens, 1999) and Kennet (McPherson,<br />

1997) to aid management decisions based on <strong>the</strong> effects<br />

of abstraction by groundwater pump<strong>in</strong>g on habitat availability.<br />

The model has also been used to assess <strong>the</strong> impact<br />

of channel restoration (Acreman and Elliott, 1996) and<br />

difference <strong>in</strong> habitat caused by different levels of channel<br />

modifi cation (Booker & Dunbar, 2004; Booker et al.,<br />

2003). Over <strong>the</strong> years, <strong>the</strong> methodology used has been<br />

adapted by various researchers and <strong>in</strong>stitutions (Table 7.4).<br />

This has lead to <strong>the</strong> development of o<strong>the</strong>r models that<br />

follow basically <strong>the</strong> same approach (Parasiewicz and<br />

Dunbar, 2001).<br />

The approach adopted <strong>in</strong> many PHABSIM studies has<br />

been outl<strong>in</strong>ed by Elliott et al. (1999) and Johnson et al.<br />

(1995). This approach <strong>in</strong>cludes identifi cation of river<br />

sectors and species of <strong>in</strong>terest, identifi cation of habitats that<br />

exist with<strong>in</strong> <strong>the</strong> river length of <strong>in</strong>terest, selection of crosssections<br />

which represent replicates of each habitat type and<br />

collection of model calibration data (water surface elevation,<br />

depth and velocity). The distribution of depths and<br />

velocities are <strong>the</strong>n predicted for <strong>the</strong> range of discharge<br />

required. Predicted depths and velocities and measured<br />

substrate classes are <strong>the</strong>n compared with HSIs. This allows


128 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Table 7.4 Examples of adaptations of PHABSIM <strong>in</strong> different countries<br />

Country <strong>River</strong> Software name Adaptation Reference<br />

France Sa<strong>in</strong>t Sauveur brook EVHA French version of PHABSIM Rousel et al. (1999)<br />

Taiwan Chou-Shui Creek PHABSIM Incorporates affects of simulated<br />

substrate changes<br />

Wu & Wang (2002)<br />

Italy Adda PHABSIM Utilises bivariate HSIs Vismara et al. (2001)<br />

USA North Fork Middle PHABSIM Incorporates an <strong>in</strong>dividual based Van W<strong>in</strong>kle et al. (1998)<br />

Fork Tule <strong>River</strong><br />

model<br />

Canada Waterton PHABSIM Uses hydraulic output from a 2D<br />

model<br />

Ghanem et al. (1996)<br />

F<strong>in</strong>land <strong>River</strong> Oulujoki FISU Uses hydraulic output from a 2D<br />

model<br />

Yrjänä et al. (1999)<br />

France Rhone STATHAB Uses a statistical hydraulics model. Lamouroux et al. (1999)<br />

Norway Mandal SSIIM Uses hydraulic output from a 3D<br />

model<br />

Fjeldstad (2001)<br />

Switzerland Brenno CASIMIR Incorporates fuzzy rules for fi sh<br />

and <strong>in</strong>vertebrate HSIs<br />

http://www.greenhydro.ch<br />

USA Qu<strong>in</strong>nebaug Meso-HABSIM Broader scale, uses more habitat<br />

variables<br />

Parasiewicz (2001)<br />

Figure 7.12 Habitat suitability <strong>in</strong>dices (HSIs) calculated for juvenile salmon preference us<strong>in</strong>g pooled results from several rivers<br />

(Source: Data from Dunbar et al., 2001)<br />

prediction of usable physical habitat for <strong>the</strong> species/life<br />

stage of <strong>in</strong>terest, as WUA <strong>in</strong> m 2 per 1000 m of river<br />

channel.<br />

The form of HSIs used <strong>in</strong> PHABSIM applications will<br />

affect <strong>the</strong> results (Booker and Dunbar, 2004) and is <strong>the</strong>refore<br />

one source of uncerta<strong>in</strong>ty affect<strong>in</strong>g results. HSIs have<br />

been categorised <strong>in</strong>to several groups. These are:<br />

• Type I: HSIs that are derived from expert op<strong>in</strong>ion or<br />

through <strong>in</strong>formation published <strong>in</strong> literature.<br />

• Type II: HSIs that have been derived through frequency<br />

analysis of physical habitat conditions used by<br />

different species or life stages as identifi ed <strong>in</strong> fi eld<br />

observations.<br />

• Type III: Type II HSIs that have been corrected for<br />

habitat availability. These HSIs are referred to as preference<br />

curves.<br />

• Type IV: Multi-variant HSIs that weight habitat suitability<br />

based on depth and velocity toge<strong>the</strong>r.<br />

Site-specifi c HSIs can be developed for particular rivers.<br />

However, this is costly and time consum<strong>in</strong>g. HSIs that<br />

used pooled results from several rivers <strong>in</strong> an attempt to<br />

create generalised curves that can be transferred between<br />

rivers are shown <strong>in</strong> Figure 7.12. Belaud et al. (1989)<br />

reported similarities between four site-specifi c HSIs and<br />

generalised HSIs, suggest<strong>in</strong>g that generalised HSIs may<br />

be more useful. Roussel et al. (1999) added to <strong>the</strong> debate


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 129<br />

on HSIs by suggested that different HSIs should be used<br />

for rest<strong>in</strong>g and feed<strong>in</strong>g fi sh. Vismara et al. (2001) compared<br />

univariate and bivariate HSIs, conclud<strong>in</strong>g that depth<br />

is much more important than velocity <strong>in</strong> defi n<strong>in</strong>g habitat<br />

suitability requirements when us<strong>in</strong>g bivariate models.<br />

Evaluation of PHABSIM<br />

The PHABSIM modell<strong>in</strong>g methodology may be applied<br />

before and after a restoration scheme to <strong>in</strong>fer changes <strong>in</strong><br />

habitat us<strong>in</strong>g a methodology such as Elliott et al. (1996).<br />

Alternatively, predicted results may be compared with<br />

changes <strong>in</strong> abundance of fi sh derived from surveys before<br />

and after <strong>the</strong> scheme. In both <strong>in</strong>stances, it is important to<br />

remember <strong>the</strong> simplicity of <strong>the</strong> PHABSIM approaches,<br />

both hydraulic and habitat modell<strong>in</strong>g. Successful application<br />

still requires considerable calibration and refi nement.<br />

Calibration may be based on empirical (regression) fi ts <strong>in</strong><br />

an effort to capture stage-dependent as well as spatial<br />

variations between vertical cross-section slices (Milhous<br />

et al., 1989). Alternatively, local and stage-dependent<br />

variations may be estimated from sparse velocity measurements<br />

at a s<strong>in</strong>gle stage by back-calculat<strong>in</strong>g of Mann<strong>in</strong>g’s<br />

n on a slice-by-slice basis, hav<strong>in</strong>g fur<strong>the</strong>r adjusted to<br />

match <strong>the</strong> modelled fl ow. In this case, however, <strong>the</strong> physical<br />

<strong>in</strong>tegrity of <strong>the</strong> model is <strong>the</strong>n lost and cells are not truly<br />

associated by sound hydraulic pr<strong>in</strong>ciples (Ghanem et al.,<br />

1996). In nei<strong>the</strong>r case, <strong>the</strong>refore, is <strong>the</strong> model a real substitute<br />

for fur<strong>the</strong>r fi eld measurement. Frequently, <strong>the</strong>re<br />

is an <strong>in</strong>teraction between <strong>the</strong> uncerta<strong>in</strong>ties <strong>in</strong>herent <strong>in</strong><br />

a one-dimensional model and those aris<strong>in</strong>g <strong>in</strong> higherorder<br />

model applications, s<strong>in</strong>ce <strong>the</strong> output of <strong>the</strong> onedimensional<br />

model is sometimes required to determ<strong>in</strong>e <strong>the</strong><br />

<strong>in</strong>itial water surface profi le for <strong>the</strong> more complex model.<br />

While PHABSIM rema<strong>in</strong>s someth<strong>in</strong>g of an ‘<strong>in</strong>dustry<br />

standard’ tool, <strong>the</strong> biological representation is also often<br />

highly simplifi ed and empirical (Pusey, 1998). Wood et al.<br />

(2001) found that <strong>the</strong> strength of ecological relationships<br />

<strong>in</strong>creased and model errors are reduced with <strong>in</strong>creas<strong>in</strong>g<br />

spatial resolution, <strong>in</strong>dicat<strong>in</strong>g that most potential is still at<br />

<strong>the</strong> site-scale but that all modell<strong>in</strong>g efforts are limited by<br />

<strong>the</strong> lack of longer-term data sets, too. However, one advantage<br />

of this habitat modell<strong>in</strong>g approach is that <strong>the</strong>re are<br />

clear manuals that defi ne step-by-step procedures, which<br />

allow replication of results by different researchers. The<br />

disadvantage of this is that it has led to poor applications<br />

by practitioners with little experience. Best results are<br />

obta<strong>in</strong>ed where teams <strong>in</strong>clud<strong>in</strong>g hydraulic eng<strong>in</strong>eers,<br />

hydrologists and ecologists work toge<strong>the</strong>r, us<strong>in</strong>g habitat<br />

modell<strong>in</strong>g as a basis for <strong>the</strong>ir river-specifi c studies. Major<br />

strengths and weaknesses of <strong>the</strong> PHABSIM methodology<br />

are summarised <strong>in</strong> Table 7.5.<br />

7.4.5 More Complex Hydrodynamic Modell<strong>in</strong>g and<br />

Habitat Simulation<br />

Numerous specifi c modell<strong>in</strong>g applications have been<br />

described that demonstrate some k<strong>in</strong>d of improvement on<br />

simpler schemes. However, <strong>the</strong>se have not given rise to<br />

any s<strong>in</strong>gle package that is <strong>the</strong> logical replacement to<br />

PHABSIM. Greater hydraulic process representation may<br />

be achieved us<strong>in</strong>g two- and three-dimensional computational<br />

fl uid dynamics models (Alfredsen et al., 1997;<br />

Booker, 2003). New approaches to quantify<strong>in</strong>g hydraulic<br />

habitat have been published (Peters et al., 1995; Nestler<br />

and Sutton, 2000). New habitat models have <strong>in</strong>cluded<br />

additional variables and have been expanded to <strong>the</strong> community<br />

level (Ba<strong>in</strong> et al., 1988; Ba<strong>in</strong>, 1995; Lamoroux<br />

et al., 1998). All of <strong>the</strong>se improvements currently come at<br />

a cost of <strong>in</strong>creased complexity, although <strong>in</strong> <strong>the</strong> future it is<br />

hoped that <strong>the</strong>y can be used to derive general rules with<br />

which to develop improved look-up methods and to defi ne<br />

<strong>the</strong> impacts of river fl ow regulation on populations ra<strong>the</strong>r<br />

than habitats (Hardy, 1998).<br />

Two-dimensional PHABSIM<br />

The PHABSIM method may also be applied us<strong>in</strong>g spatially<br />

cont<strong>in</strong>uous patterns of depth and velocity as predicted<br />

by hydrodynamic models (e.g. Ghanem et al.,<br />

1996). Spatially cont<strong>in</strong>uous <strong>in</strong>formation negates any problems<br />

caused by sampl<strong>in</strong>g frequencies applied <strong>in</strong> fi eld<br />

based studies. This approach also allows visualisation of<br />

areas of suitable habitat. A comparison of habitat suitability<br />

predicted us<strong>in</strong>g two different HSIs for a 50 m reach<br />

of <strong>the</strong> Bere stream, Dorset, UK, is shown <strong>in</strong> Figure 7.13.<br />

These maps show how truncation of <strong>the</strong> depth HSI can<br />

lead to changes <strong>in</strong> patterns of habitat suitability. An<br />

accompany<strong>in</strong>g sensitivity analysis dur<strong>in</strong>g <strong>the</strong> construction<br />

of Figure 7.13 also demonstrates how uncerta<strong>in</strong>ty <strong>in</strong><br />

habitat modell<strong>in</strong>g will depend on <strong>the</strong> <strong>in</strong>teractions between<br />

<strong>the</strong> sampl<strong>in</strong>g density and <strong>the</strong> HSI used. In analogy with<br />

<strong>the</strong> analysis of velocity data <strong>in</strong> Figure 7.9, spatially cont<strong>in</strong>uous<br />

simulations can be sub-sampled to simulate how<br />

<strong>the</strong> number of cross-sections and po<strong>in</strong>t measurements<br />

might affect results had <strong>the</strong> <strong>in</strong>formation been collected <strong>in</strong><br />

<strong>the</strong> fi eld. In general, predicted habitat decreases as more<br />

po<strong>in</strong>ts are added across each section. This occurs because<br />

poor habitat at <strong>the</strong> marg<strong>in</strong>s of <strong>the</strong> channel is more likely<br />

to be sampled when more po<strong>in</strong>ts are measured along each<br />

section. When <strong>the</strong> truncated HSIs were used, <strong>the</strong> positions<br />

of <strong>the</strong> cross-sections <strong>in</strong> relation to a deep pool <strong>in</strong> <strong>the</strong><br />

middle of <strong>the</strong> reach had a strong <strong>in</strong>fl uence on predicted<br />

habitat. When small numbers of cross-sections were used,<br />

<strong>the</strong> likelihood of <strong>the</strong> pool be<strong>in</strong>g over- or under-sampled


130 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Table 7.5 Major strengths and weaknesses of <strong>the</strong> application of <strong>the</strong> PHABSIM methodology<br />

Strengths Weaknesses<br />

The method used is replicable and <strong>the</strong>refore<br />

does not rely on expert op<strong>in</strong>ion<br />

The method can be applied us<strong>in</strong>g hydraulic<br />

predictions calculated us<strong>in</strong>g a range of<br />

techniques from crude stage-discharge<br />

relationships to spatially cont<strong>in</strong>uous<br />

computational fl uid dynamics modell<strong>in</strong>g.<br />

Channel change can be <strong>in</strong>corporated through<br />

repeated survey<br />

Multi-variate habitat suitability <strong>in</strong>dices have<br />

been developed (Vismara et al., 2001).<br />

Application of habitat mapp<strong>in</strong>g has allowed<br />

results to be up-scaled to represent <strong>the</strong><br />

length of river under <strong>in</strong>vestigation<br />

(Maddock and Bird, 1996; Maddock, 1999).<br />

Recent studies have confi rmed that physical<br />

habitat is an important factor controll<strong>in</strong>g<br />

fi sh populations <strong>in</strong> <strong>the</strong> long term (Sabaton<br />

et al., 2003; Souchon and Capra, 2003).<br />

No readily applicable alternative methods<br />

have been developed which enable<br />

prediction of population changes as <strong>the</strong><br />

result of water resource impacts.<br />

was higher. Results also suggest that, as <strong>the</strong> number of<br />

cross-section <strong>in</strong>creases, <strong>the</strong>re is less variation <strong>in</strong> predicted<br />

habitat above six cross-sections when <strong>the</strong> non-truncated<br />

HSIs were used and eight cross-sections when <strong>the</strong> truncated<br />

HSIs were used.<br />

Swimm<strong>in</strong>g Speeds<br />

A more sophisticated example of habitat modell<strong>in</strong>g us<strong>in</strong>g<br />

hydrodynamic calculations is to compare patterns of<br />

velocities predicted by three-dimensional models with <strong>the</strong><br />

swimm<strong>in</strong>g speeds of fi sh to assess <strong>the</strong> effects of channel<br />

design on fi sh habitat. The swimm<strong>in</strong>g performance of fi sh<br />

can also be analysed to assess <strong>the</strong> health of <strong>the</strong> fi sh when<br />

exposed to sub-lethal toxic chemicals (e.g. Alsop et al.,<br />

1999). One frequently used measurement is <strong>the</strong> Maximum<br />

Results may be sensitive to <strong>the</strong> location of cross-sections and<br />

<strong>the</strong> position of measurements along <strong>the</strong>se cross-sections<br />

Inaccuracies <strong>in</strong> hydraulic modell<strong>in</strong>g, especially <strong>in</strong> steep boulder<br />

rivers (Azzell<strong>in</strong>o & Vismara, 2001) and where vegetation<br />

growth causes changes <strong>in</strong> stage-discharge relationships<br />

throughout <strong>the</strong> year (Hearne et al. 1994) will affect habitat<br />

predictions.<br />

Requires re-calibration to assess changes <strong>in</strong> channel morphology<br />

over time. Incorporation of channel change would require<br />

expensive repeated surveys or sediment transport predictions,<br />

which have <strong>the</strong>ir own sources of uncerta<strong>in</strong>ty.<br />

Rules for weight<strong>in</strong>g depth, velocity and substrate are not readily<br />

available.<br />

Typically restricted to depth, velocity and substrate, although<br />

variables such as water quality, vegetation cover can be<br />

<strong>in</strong>corporated.<br />

Reached-based results (cover<strong>in</strong>g approximately 100 m of river<br />

length) must be scaled up.<br />

Calculates habitat suitability and not population numbers or<br />

presence/absence.<br />

In most studies physical habitat availability at low fl ow has been<br />

<strong>the</strong> ma<strong>in</strong> area of <strong>in</strong>terest, with PHABSIM used primarily to<br />

compare <strong>the</strong> implications of alternative fl ow regulation<br />

scenarios on habitat. This has led to an emphasis on <strong>the</strong><br />

relationship between discharge and usable habitat given a<br />

distribution of relatively shallow depths and slow velocities.<br />

Susta<strong>in</strong>ed Swimm<strong>in</strong>g Speed (MSSS). This is defi ned<br />

as <strong>the</strong> maximum velocity at which a fi sh can swim for<br />

a period of more than 200 m<strong>in</strong>utes (Turnpenny et al.,<br />

2001).<br />

Booker (2003) compared simulated velocity patterns to<br />

MSSS of roach, dace and chub to assess habitat suitability<br />

at high fl ows <strong>in</strong> two reaches of <strong>the</strong> <strong>River</strong> Tame, an urban<br />

river <strong>in</strong> Birm<strong>in</strong>gham, UK. The percentage of <strong>in</strong>-stream<br />

area that was less than <strong>the</strong> MSSS was used as an <strong>in</strong>dicator<br />

of habitat quality. This was <strong>the</strong> area of river, at a specifi ed<br />

distance from <strong>the</strong> bed, <strong>in</strong> which fi sh of a certa<strong>in</strong> size and<br />

species could <strong>the</strong>oretically susta<strong>in</strong> a position for at least<br />

200 m<strong>in</strong>utes. Due to uncerta<strong>in</strong>ty as to <strong>the</strong> exact distance<br />

between <strong>the</strong> fi sh and <strong>the</strong> river bed dur<strong>in</strong>g high fl ows, and<br />

to simplify <strong>the</strong> three-dimensional nature of <strong>the</strong> analysis,<br />

different heights above <strong>the</strong> bed were considered separately.


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 131<br />

Figure 7.13 Maps of habitat suitability derived us<strong>in</strong>g <strong>the</strong> same hydraulic patterns but different suitability curves: (a) HSIs given <strong>in</strong><br />

Dunbar et al. (2001), (b) <strong>the</strong> same HSIs truncated for depth


132 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

The percentage of survivable area for chub, dace and<br />

roach at bankfull discharge was calculated at two sites<br />

(Figure 7.14). The ‘Highly Modifi ed’ reach was a 91 m<br />

straightened reach with an average width of 11.9 m and<br />

artifi cially streng<strong>the</strong>ned banks conta<strong>in</strong>ed with<strong>in</strong> a larger<br />

two-stage channel. This reach had no dist<strong>in</strong>ct geomorphological<br />

features and relatively uniform bed topography. In<br />

contrast, <strong>the</strong> ‘Less Modifi ed’ reach was 139 m <strong>in</strong> length<br />

with an average width of 10.0 m. This reach was also<br />

conta<strong>in</strong>ed with<strong>in</strong> a two-stage channel but had one bank<br />

consist<strong>in</strong>g of natural material, a slightly s<strong>in</strong>uous path and<br />

an undulat<strong>in</strong>g bed profi le. Figure 7.14(b) shows that as <strong>the</strong><br />

body length <strong>in</strong>creases <strong>the</strong>re is an <strong>in</strong>crease <strong>in</strong> <strong>the</strong> area of<br />

habitat <strong>in</strong> which a fi sh is likely to be able to susta<strong>in</strong> a stationary<br />

position. This is because bigger fi sh can susta<strong>in</strong><br />

faster swimm<strong>in</strong>g speeds. Similarly, as distance from <strong>the</strong><br />

bed <strong>in</strong>creases <strong>the</strong> percentage of survivable habitat<br />

decreases due to faster velocities. Results showed that at<br />

<strong>the</strong> Highly Modifi ed site <strong>the</strong> uniform channel structure<br />

supported no discrete areas of slower velocity which could<br />

be used as refugia. At <strong>the</strong> Less Modifi ed site a deep pool<br />

approximately 40 m from <strong>the</strong> upstream boundary provided<br />

a fl ow refuge that could act as a niche of suitable habitat.<br />

There was also a separate discrete area of slower velocity<br />

created by <strong>the</strong> shelter<strong>in</strong>g effect of a change <strong>in</strong> direction of<br />

<strong>the</strong> river planform.<br />

Bioenergetics<br />

Bioenergetic models attempt to quantify <strong>the</strong> trade-off<br />

between energy ga<strong>in</strong>ed through feed<strong>in</strong>g and that lost<br />

through swimm<strong>in</strong>g ‘costs’. Costs arise from hold<strong>in</strong>g position<br />

<strong>in</strong> a mov<strong>in</strong>g fl ow, <strong>in</strong> <strong>the</strong> capture of food, as well as<br />

through digestion, faeces and ur<strong>in</strong>e (Hayes et al., 2000).<br />

Fish activity may be directly observed and related to fl ow<br />

conditions or estimated on <strong>the</strong> basis of models that use<br />

results of physiological experiments undertaken on fi sh<br />

which are forced to swim aga<strong>in</strong>st fl ows of constant<br />

velocity. For example, Booker et al. (2004b) used a threedimensional<br />

Computational Fluid Dynamics (3D-CFD)<br />

model to simulate hydraulic patterns <strong>in</strong> a 50 m reach of<br />

<strong>the</strong> Bere stream, Dorset, UK. This <strong>in</strong>formation was <strong>the</strong>n<br />

comb<strong>in</strong>ed with a bioenergetic model that used behavioural<br />

and physiological relationships to quantify <strong>the</strong> spatial<br />

pattern of energy ga<strong>in</strong> when feed<strong>in</strong>g on <strong>in</strong>vertebrates drift<strong>in</strong>g<br />

<strong>in</strong> <strong>the</strong> river. The model was tested by compar<strong>in</strong>g patterns<br />

of predicted energy <strong>in</strong>take with observed habitat use<br />

by juvenile salmonids at different times of day. A map of<br />

energy <strong>in</strong>take predicted by <strong>the</strong> model compared with<br />

observed locations of feed<strong>in</strong>g and rest<strong>in</strong>g fi sh at <strong>the</strong> site<br />

is shown <strong>in</strong> Figure 7.15.<br />

There are many uncerta<strong>in</strong>ties that must be considered<br />

when us<strong>in</strong>g this bioenergetic modell<strong>in</strong>g approach as a tool<br />

to assist river restoration. The approach relies on <strong>the</strong> accuracy<br />

of both <strong>the</strong> hydrodynamic predictions and <strong>the</strong> accuracy<br />

of <strong>the</strong> physiological algorithms and parameters used.<br />

Enders et al. (2003) demonstrate that models of fi sh activity<br />

cost based upon constant velocity experiments may<br />

underestimate actual costs by up to a factor of 4.2, because<br />

turbulent fl ow fl uctuations are neglected. Direct observations<br />

of feed<strong>in</strong>g behaviour illustrate close relationships<br />

to fi sh activity and characteristics of turbulent fl ow (Enders<br />

Figure 7.14 (a) Mean ‘maximum susta<strong>in</strong>able swimm<strong>in</strong>g speed’ and ‘burst swimm<strong>in</strong>g speed’ for roach, dace and chub at 8 °C (based<br />

on data from Clough & Turnpenny, 2001), (b) Percentage volume of habitat less than <strong>the</strong> mean ‘maximum susta<strong>in</strong>able swimm<strong>in</strong>g<br />

speed’ at different distances from <strong>the</strong> bed <strong>in</strong> each reach at 18 m 3 s −1 (Source: Figure 11 of Booker, 2003)


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 133<br />

Figure 7.15 Predicted net energy <strong>in</strong>take for a 0.1 m fi sh and observed fi sh locations<br />

et al., 2005). Models may also be hard to test aga<strong>in</strong>st fi eld<br />

observations of habitat uses because factors o<strong>the</strong>r than<br />

energy <strong>in</strong>take can <strong>in</strong>fl uence microhabitat selection. For<br />

example, <strong>the</strong> proximity to o<strong>the</strong>r fi sh (Valdimarsson and<br />

Metcalfe, 2001) and predation risk or distance to cover<br />

(Mesick, 1988). Fish do not spend 100% of <strong>the</strong>ir time driftfeed<strong>in</strong>g.<br />

They may feed very effi ciently for short periods<br />

and <strong>the</strong>n retreat to more sheltered locations (Gries and<br />

Juanes, 1998). Fur<strong>the</strong>rmore, <strong>the</strong> decision to select a certa<strong>in</strong><br />

position may not result from conditions (e.g. drift density)<br />

at that time but conditions at a previous time or conditions<br />

over a longer period. Also, a fi sh may not have perfect<br />

knowledge of its habitat. This means that fi sh feed<strong>in</strong>g <strong>in</strong> an<br />

energetically poor area may not be aware that <strong>the</strong>re are<br />

more favourable alternative positions elsewhere.<br />

7.4.6 Alternatives to Physical Habitat Models<br />

Several alternatives to complex eco-hydraulic simulation<br />

have emerged. A fi eld approach known as ‘Expert Habitat<br />

Mapp<strong>in</strong>g’ (EHM) is used <strong>in</strong> streams that are hydraulically<br />

complex, particularly where <strong>the</strong> irregularity and variability<br />

of boundary conditions makes hydraulic modell<strong>in</strong>g<br />

<strong>in</strong>appropriate (for example, <strong>in</strong> steep bedrock rivers with<br />

large woody debris). EHM uses habitat suitability criteria<br />

but <strong>the</strong>n relies on fi sh biologists mapp<strong>in</strong>g habitat based on<br />

those criteria at a range of fl ows to develop fl ow-vs-habitat<br />

curves. The method works well provided that <strong>the</strong> HSI cri-<br />

teria are used and verifi ed with spot measurements, and<br />

<strong>the</strong> biologists are experienced.<br />

Alternatively, physically-parameterised models may be<br />

replaced by stochastic or hybrid approaches such as cellular<br />

autometa (Chen et al., 2002) or neural network<br />

models (Werner and Oback, 2001; Reyjol et al., 2001;<br />

Gevrey et al., 2003) as witnessed <strong>in</strong> hydrology and fl oodpla<strong>in</strong><br />

<strong>in</strong>undation studies. As yet, however, <strong>the</strong>se approaches<br />

have focused on simulat<strong>in</strong>g ecological characteristics:<br />

appropriate scales for applications, for <strong>the</strong> development of<br />

rules and tra<strong>in</strong><strong>in</strong>g data sets rema<strong>in</strong> to be explored.<br />

7.5 CONCLUSIONS<br />

Restor<strong>in</strong>g and rehabilitat<strong>in</strong>g rivers for ecological purposes<br />

is an essentially multi-discipl<strong>in</strong>ary and multi-stage activity.<br />

There is grow<strong>in</strong>g awareness that to be successful (that<br />

is, to produce schemes which are susta<strong>in</strong>able <strong>in</strong> <strong>the</strong><br />

medium- to longer-term) truly functional environments<br />

are required, where catchment hydrology provides <strong>the</strong><br />

background, or context, for reach and <strong>in</strong>ter-reach fl ows<br />

and sediment transport. In turn, <strong>the</strong>se hydraulic variables<br />

structure physical habitats, which <strong>the</strong>mselves support a<br />

diverse ecological response. Monitor<strong>in</strong>g of fl ows and<br />

aquatic ecology have been long-stand<strong>in</strong>g areas of research<br />

and cont<strong>in</strong>ue to be vitally important but, <strong>in</strong>creas<strong>in</strong>gly,<br />

emphasis has been upon <strong>the</strong> modell<strong>in</strong>g or simulation of<br />

fl ows and <strong>the</strong> habitats which <strong>the</strong>se determ<strong>in</strong>e. Modell<strong>in</strong>g


134 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

is useful <strong>in</strong> extend<strong>in</strong>g <strong>the</strong> range of fi eld observation, <strong>in</strong><br />

scenario-type exercises to evaluate restoration schemes<br />

before <strong>the</strong>ir implementation and also <strong>in</strong> provid<strong>in</strong>g outputs<br />

aga<strong>in</strong>st which schemes might be appraised after implementation.<br />

Ideally, modell<strong>in</strong>g efforts should proceed with<br />

<strong>the</strong> closest possible coupl<strong>in</strong>g between <strong>the</strong>ir physical<br />

(hydraulic and hydrodynamic) and biological elements,<br />

and <strong>in</strong> close association with improved strategies for fi eld<br />

monitor<strong>in</strong>g, too. Runn<strong>in</strong>g through this chapter are some<br />

very basic <strong>the</strong>mes. These refl ect <strong>the</strong> pr<strong>in</strong>cipal sources of<br />

uncerta<strong>in</strong>ty <strong>in</strong> restoration schemes for ecological purposes<br />

and must be addressed to improve both fi eld and modell<strong>in</strong>g<br />

aspects of restoration design:<br />

data requirements – <strong>the</strong> appropriate amount and k<strong>in</strong>d of<br />

fi eld data required to adequately specify pre- and post<br />

design aspects of <strong>the</strong> river;<br />

characterisation – of fl ow coherence and species assemblages<br />

or behaviour, which partly follows from data<br />

considerations as above, but which also <strong>in</strong>cludes design<br />

criteria relat<strong>in</strong>g to channel form, as well as parameterisation<br />

of models;<br />

coupl<strong>in</strong>g – of key physical and ecological aspects of river<br />

system function <strong>in</strong> models;<br />

awareness – of model application, limitations and sensitivity,<br />

and of o<strong>the</strong>r limitations;<br />

development – of newer, more fl exible means of model<br />

evaluation and of restoration appraisal more generally.<br />

While each of <strong>the</strong> above rema<strong>in</strong> sources of uncerta<strong>in</strong>ty,<br />

<strong>the</strong>y represent, too, areas of opportunity for <strong>the</strong> rapid<br />

development of research and <strong>in</strong>tervention protocols. For<br />

<strong>the</strong> practitioner, ask<strong>in</strong>g questions from each of <strong>the</strong>se areas<br />

– and questions which cut across or l<strong>in</strong>k between <strong>the</strong>m – is<br />

perhaps <strong>the</strong> best form of guidance for recognis<strong>in</strong>g, reduc<strong>in</strong>g<br />

and even <strong>in</strong>corporat<strong>in</strong>g uncerta<strong>in</strong>ty <strong>in</strong>to river restoration<br />

activity from an ecological standpo<strong>in</strong>t.<br />

ACKNOWLEDGEMENTS<br />

This chapter was produced as part of research projects<br />

NER/A/S/1998/00009, ‘Identifi cation of physically-based<br />

design criteria for riffl e-pool sequences <strong>in</strong> river rehabilitation’,<br />

NER/D/S/2000/01422, ‘Formation of a new river<br />

channel: fl ow, sediment and vegetation dynamics’, and<br />

NER/T/S/2001/01250, ‘Vegetation <strong>in</strong>fl uences on fi ne sediment<br />

and propagule dynamics <strong>in</strong> groundwater-fed rivers:<br />

implications for river management, restoration and riparian<br />

biodiversity’. These were funded <strong>in</strong> <strong>the</strong> United<br />

K<strong>in</strong>gdom by <strong>the</strong> Natural Environment Research Council.<br />

SSIIM is produced and made available by Nils R. B.<br />

Olsen, Department of Hydraulic and Environmental<br />

Eng<strong>in</strong>eer<strong>in</strong>g, The Norwegian University of Science and<br />

Technology. Scott McBa<strong>in</strong> k<strong>in</strong>dly provided references to<br />

<strong>the</strong> fi gures and tabular <strong>in</strong>formation used <strong>in</strong> <strong>the</strong> text, and<br />

advised on <strong>in</strong>itial chapter layout and content. Nigel Wright<br />

made additional suggestions <strong>in</strong> respect of hydraulic<br />

modell<strong>in</strong>g.<br />

REFERENCES<br />

Acreman MC. 2000. The chang<strong>in</strong>g hydrology of <strong>the</strong> UK. In:<br />

Acreman MC (Ed), The Hydrology of <strong>the</strong> UK: A Study of<br />

Change. Routledge: London, UK.<br />

Acreman MC. 2002. Case Studies of Managed Flood Releases.<br />

Environmental Flow Assessment Part III. World Bank Water<br />

Resources and Environmental Management Best Practice Brief<br />

No 8, World Bank: Wash<strong>in</strong>gton, DC.<br />

Acreman MC, Elliott CRN. 1996. Evaluation of <strong>the</strong> river Wey<br />

restoration project us<strong>in</strong>g <strong>the</strong> Physical HABitat SIMuation<br />

(PHABSIM) model. Proceed<strong>in</strong>gs of <strong>the</strong> 31st MAFF Conference<br />

of <strong>River</strong> and Coastal Eng<strong>in</strong>eers, Keele University, 3–5<br />

July 1996; 1.2.1–1.2.14.<br />

Acreman MC, K<strong>in</strong>g J. 2003. Defi n<strong>in</strong>g water requirements. In:<br />

Dyson M, Bergkamp G, Scanlon H (Eds), Flow: The Essentials<br />

of Environmental Flows. IUCN: Gland, Switzerland and Cambridge,<br />

UK.<br />

Alfredsen K, Marchand W, Bakken TH, Harby A. 1997. Application<br />

and comparison of computer models quantify<strong>in</strong>g impacts<br />

of river regulation on fi sh habitat. In: Broch E, Lysne DK,<br />

Flatabo N, Helland-Hansen E (Eds), Proceed<strong>in</strong>gs of <strong>the</strong> 3rd<br />

International conference on hydropower Hydropower ’97, A.A.<br />

Balkema Publishers: Rotterdam/Brookfi eld.<br />

Alsop DH, McGeer JC, McDonald DG, Wood CM. 1999. Costs<br />

of chronic waterborne z<strong>in</strong>c exposure and <strong>the</strong> consequences of<br />

z<strong>in</strong>c acclimation on <strong>the</strong> gill/z<strong>in</strong>c <strong>in</strong>teractions of ra<strong>in</strong>bow trout<br />

<strong>in</strong> hard and soft water. Environmental Toxicology and Chemistry<br />

18: 1014–1025.<br />

Anderson MG, Wall<strong>in</strong>g DE, Bates PD (Eds). 1996. Floodpla<strong>in</strong><br />

Processes. John Wiley & Sons Ltd: Chichester, UK.<br />

Archer D, Newson, M. 2002. The use of <strong>in</strong>dices of fl ow variability<br />

<strong>in</strong> assess<strong>in</strong>g <strong>the</strong> hydrological and <strong>in</strong>stream habitat impacts of<br />

upland afforestation and dra<strong>in</strong>age. Journal of Hydrology 268:<br />

244–258.<br />

Armitage PD, Petts GE. 1992. Biotic score and prediction to<br />

assess <strong>the</strong> effects of water abstraction on river macro<strong>in</strong>vertebrates<br />

for conservation purposes. Aquatic Conservation:<br />

Mar<strong>in</strong>e and Freshwater Ecosystems 2: 1–17.<br />

Arth<strong>in</strong>gton AH, Long GC (Eds). 1997. Logan <strong>River</strong> Trial of <strong>the</strong><br />

Build<strong>in</strong>g Block Methodology for Assess<strong>in</strong>g Environmental<br />

Flow Requirements: Background Papers. Centre for Catchment<br />

and In-Stream Research and Department of Natural Resoruces:<br />

Brisbane, Queensland.<br />

Arth<strong>in</strong>gton AH, Lloyd R (Eds). 1998. Logan <strong>River</strong> Trial of <strong>the</strong><br />

Build<strong>in</strong>g Block Methodology for Assess<strong>in</strong>g Environmental<br />

Flow Requirements: Workshop Report. Centre for Catchment<br />

and In-Stream Research and Dept Natural Resources: Brisbane,<br />

Queensland, Australia.


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 135<br />

Azzell<strong>in</strong>o A, Vismara R. 2001. Pool quality <strong>in</strong>dex: New method<br />

to defi ne m<strong>in</strong>imum fl ow requirements of high-gradient, loworder<br />

streams. Journal of Environmental Eng<strong>in</strong>eer<strong>in</strong>g 127 (11):<br />

1003–1013.<br />

Barker I, Kirmond A. 1998. <strong>Manag<strong>in</strong>g</strong> surface water abstraction.<br />

In: Wheater H, Kirby C (Eds), Hydrology <strong>in</strong> a Chang<strong>in</strong>g Environment<br />

British Hydrological Society, 249–258.<br />

Ba<strong>in</strong> MB. 1995. Habitat at <strong>the</strong> local scale: multivariate patterns for<br />

stream fi shes. Bull. Fr. Peche Pisic. 337/338/339: 165–177.<br />

Ba<strong>in</strong> MB, F<strong>in</strong>n JT, Booke HE. 1988. Streamfl ow regulation and<br />

fi sh community structure. Ecology 69: 382–392.<br />

Belaud A, Chaveroche P, Lim P, Sabaton C. 1989. Probability-ofuse<br />

curves applied to brown trout (Salmo trutta fario L.) <strong>in</strong><br />

rivers of sou<strong>the</strong>rn France. Regulated <strong>River</strong>s: Research and<br />

Management 3: 321–336.<br />

Benetti AD, Lanna AE, Cobalch<strong>in</strong>i MS. 2002. Current practice<br />

for establish<strong>in</strong>g environmental fl ows <strong>in</strong> Brazil. Proceed<strong>in</strong>gs of<br />

<strong>the</strong> Fourth International Ecohydraulics Symposium, Sou<strong>the</strong>rn<br />

Waters, Cape Town.<br />

Bergen SD, Bolton SM, Fridley JL. 2001. Design Pr<strong>in</strong>ciples for<br />

ecological eng<strong>in</strong>eer<strong>in</strong>g. Ecological Eng<strong>in</strong>eer<strong>in</strong>g 18: 201–210.<br />

Beven K. 2000. Ra<strong>in</strong>fall–runoff Modell<strong>in</strong>g. John Wiley & Sons<br />

Ltd: Chichester, UK.<br />

Beven K. 2001. On modell<strong>in</strong>g as collective <strong>in</strong>telligence. Hydrological<br />

Processes 15: 2205–2207.<br />

Bierkens MFP, F<strong>in</strong>ke PA, de Willigen P. 2000. Upscal<strong>in</strong>g and<br />

Downscal<strong>in</strong>g Methods for Environmental Research. Kluwer<br />

Academic Publishers: Dordrecht, The Ne<strong>the</strong>rlands.<br />

Booker DJ. 2003. Hydraulic modell<strong>in</strong>g of fi sh habitat <strong>in</strong> urban<br />

rivers dur<strong>in</strong>g high fl ows. Hydrological Processes 17: 577–599.<br />

Booker DJ, Dunbar MJ. 2004. Application of Physical HAbitat<br />

SIMulation (PHABSIM) modell<strong>in</strong>g to modifi ed urban river<br />

channels. <strong>River</strong> Research and Applications 20: 167–183.<br />

Booker DJ, Dunbar MJ, Acreman MC et al. 2004a. Habitat assessment<br />

at <strong>the</strong> catchment scale; application to <strong>the</strong> <strong>River</strong> Itchen,<br />

UK. British Hydrological Society International Conference:<br />

Hydrology: Science & Practice for <strong>the</strong> 21st Century, July 2004,<br />

volume II, 10–18, British Hydrological Society, London, UK.<br />

Booker DJ, Dunbar MJ, Ibbotson AT. 2004b Predict<strong>in</strong>g juvenile<br />

salmonid drift-feed<strong>in</strong>g habitat quality us<strong>in</strong>g a three-dimensional<br />

hydraulic-bioenergetic model. Ecological Modell<strong>in</strong>g<br />

177: 157–177.<br />

Booker DJ, Dunbar MJ, Shamseld<strong>in</strong> A et al. 2003. Physical<br />

habitat assessment <strong>in</strong> urban rivers under future fl ow scenarios.<br />

CIWEM Journal 17: 251–257.<br />

Boon PJ. 1998. <strong>River</strong> restoration <strong>in</strong> fi ve dimensions. Aquatic Conservation:<br />

Mar<strong>in</strong>e and Freshwater Ecosystems 8: 257–264.<br />

Bovee KD. 1982. A Guide to Stream Habitat Analysis us<strong>in</strong>g <strong>the</strong><br />

IFIM. US Fish and Wildlife Service Report FWS/OBS-82/26:<br />

Fort Coll<strong>in</strong>s, Colorado.<br />

Bovee KD, Lamb BL, Bartholow JM et al. 1998. Stream habitat<br />

analysis us<strong>in</strong>g <strong>the</strong> Instream Flow Incremental Methodology. US<br />

Geological Survey, Biological Resources Division, Information<br />

and Technology Report USGS/BRD/ITR–1998–004.<br />

Bunn SE, Arth<strong>in</strong>gton AH. 2002. Basic pr<strong>in</strong>ciples and ecological<br />

consequences of altered fl ow regimes for aquatic biodiversity.<br />

Environmental Management 50: 492–507.<br />

Chen Q, Mynett AE, M<strong>in</strong>ns AW. 2002. Application of cellular<br />

automata to modell<strong>in</strong>g competitive growths of two underwater<br />

specieis Chara aspera and Potamogeton pect<strong>in</strong>atus <strong>in</strong> Lake<br />

Veluwe. Ecological Modell<strong>in</strong>g 147: 253–265.<br />

Chow VT. 1973. Open Channel Hydraulics. McGraw-Hill: Tokyo,<br />

Japan.<br />

Clarke SJ, Bruce-Burgess L, Wharton G. 2003. L<strong>in</strong>k<strong>in</strong>g form and<br />

function: towards an eco-hydromorphic approach to susta<strong>in</strong>able<br />

rive restoration. Aquatic Conservation: Mar<strong>in</strong>e and Freshwater<br />

Ecosystems 13: 439–450.<br />

Clausen B, Biggs BJF. 2000. Flow variables for ecological studies<br />

<strong>in</strong> temperate streams: group<strong>in</strong>gs based on covariance. Journal<br />

of Hydrology 237: 184–197.<br />

Clifford NJ. 2001. Conservation and <strong>the</strong> river channel environment.<br />

In: Warren A, French JR (Eds), Habitat Conservation.<br />

John Wiley & Sons Ltd: Chichester, UK; 67–104.<br />

Clifford NJ. 2002. Hydrology: <strong>the</strong> chang<strong>in</strong>g paradigm. Progress<br />

<strong>in</strong> Physical Geography 26: 290–301.<br />

Clifford NJ, Soar PJ, Emery JC et al. 2002a. Numerical fl ow modell<strong>in</strong>g<br />

for eco-hydraulic and river rehabilitation applications:<br />

a case study of <strong>the</strong> <strong>River</strong> Cole, Birm<strong>in</strong>gham, UK. In: Bousmar<br />

D, Zech Y (Eds), <strong>River</strong> Flow 2002. Lisse: Swets & Zeitl<strong>in</strong>ger/<br />

Balkema: Rotterdam, The Ne<strong>the</strong>rlands; 1195–1204.<br />

Clifford NJ, Soar PJ, Emery JC et al. 2002b. Susta<strong>in</strong><strong>in</strong>g water<br />

related ecosystems – <strong>the</strong> role of <strong>in</strong>-stream bedform design <strong>in</strong><br />

river channel rehabilitation. In: Friend 2002 – Regional hydrology:<br />

bridg<strong>in</strong>g <strong>the</strong> gap between research and practice. IAHS<br />

Publication 274: 407–413, IAHS, Wall<strong>in</strong>gford, UK.<br />

Clifford NJ, Soar PJ, Harmar OP et al. 2005. Assessment of<br />

hydrodynamic simulation results for eco-hydrological and ecohydraulic<br />

applications: a spatial semivariance approach.<br />

Hydrological Processes 19: 3631–3648.<br />

Clifford NJ, Harmar OP, Harvey G, Petts GE. 2006. Physical<br />

habitat, eco-hydraulics and river design: a review and reevaluation<br />

of some popular concepts and methods. Aquatic Conservation:<br />

Mar<strong>in</strong>e and Freshwater Ecosystems 16: 389–408.<br />

Clough SC, Turnpenny AWH. 2001. Swimm<strong>in</strong>g Speeds <strong>in</strong> Fish:<br />

Phase 1. Environment Agency R&D Technical Report W2–<br />

026/TR1, Environment Agency: Bristol, UK.<br />

Crowder DW, Diplas P. 2000. Us<strong>in</strong>g two-dimensional hydrodynamic<br />

models at scales of ecological importance. Journal of<br />

Hydrology 230: 172–191.<br />

Dakova S, Uzunov Y, Mandadjiev D. 2000. Low fl ow – <strong>the</strong> river’s<br />

ecosystem limit<strong>in</strong>g factor. Ecological Eng<strong>in</strong>eer<strong>in</strong>g 16: 167–<br />

174.<br />

Dunbar MJ, Ibbotson AT, Gow<strong>in</strong>g IM et al. 2001. Ecologically<br />

Acceptable Flows Phase III: Fur<strong>the</strong>r Validation of PHABSIM<br />

for <strong>the</strong> Habitat Requirements of Salmonid Fish. F<strong>in</strong>al R&D<br />

Technical report to <strong>the</strong> Environment Agency, Environment<br />

Agency, Bristol, UK.<br />

Dyson M, Bergkamp G, Scanlon H (Eds). 2003 Flow: The Essentials<br />

of Environmental Flows. IUCN: Gland, Switzerland and<br />

Cambridge, UK.<br />

Elliott CRN, Johnson IW, Sekul<strong>in</strong> AE et al. 1996. Guide to <strong>the</strong><br />

use of <strong>the</strong> Physical Habitat Simulation System, EA Release<br />

version. Environment Agency R&D Technical Report W20.<br />

Environment Agency: Bristol, UK.


136 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Elliott CRN, Dunbar MJ, Gow<strong>in</strong>g I, Acreman MC. 1999. A<br />

habitat assessment approach to <strong>the</strong> management of groundwater<br />

dom<strong>in</strong>ated rivers, Hydrological Processes 13: 459–475.<br />

Emery JC. 2003. Characteristics and controls of gravel-bed riffl epool<br />

sequences for habitat assessment and river rehabilitation<br />

design. Unpublished PhD <strong>the</strong>sis, University of Birm<strong>in</strong>gham,<br />

UK.<br />

Emery JC, Gurnell AM, Clifford NJ et al. 2003. Classify<strong>in</strong>g <strong>the</strong><br />

performance of riffl e-pool bedforms for habitat assessment and<br />

river rehabilitation design. <strong>River</strong> Research and Applications<br />

19: 533–549.<br />

Enders EC, Boisclair D, Roy AG. 2003. The effect of turbulence<br />

on <strong>the</strong> cost of swimm<strong>in</strong>g for juvenile Atlantic salmon (Salmos<br />

salar). Canadian Journal of Fisheries and Aquatic Sciences 60:<br />

1149–1160.<br />

Enders EC, Buffi ngton-Belanger T, Boisclair D, Roy AG. 2005.<br />

The feed<strong>in</strong>g behaviour of juvenile Atlantic salmon <strong>in</strong> relation<br />

to turbulent fl ow. Journal of Fish Biology 66: 242–253.<br />

Everard M. 1996. The importance of periodic droughts for ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g<br />

diversity <strong>in</strong> <strong>the</strong> freshwater environment. Freshwater<br />

Forum 7: 33–50.<br />

Extence C, Balbi DM, Chadd RP. 1999. <strong>River</strong> fl ow <strong>in</strong>dex<strong>in</strong>g us<strong>in</strong>g<br />

British benthic macro-<strong>in</strong>vertebrates: a framework for sett<strong>in</strong>g<br />

hydro-ecological objectives. Regulated <strong>River</strong>s Research and<br />

Management 15: 543–574.<br />

Frissell CA, Liss WJ, Warren CE, Hurley MD. 1986. A hierarchical<br />

framework for stream classifi cation: view<strong>in</strong>g streams <strong>in</strong> a<br />

watershed context. Environmental Management 10: 199–214.<br />

French JR, Clifford NJ. 2000. Hydrodynamic modell<strong>in</strong>g as a basis<br />

for expla<strong>in</strong><strong>in</strong>g estuar<strong>in</strong>e environmental dynamics: some conceptual<br />

and methodological issues. Hydrological Processes 14:<br />

2089–2108.<br />

Fjeldstad H-P. 2001. Numerical modell<strong>in</strong>g tools for predict<strong>in</strong>g<br />

physical habitat adjustments. In: Taugbol T, L’Abee-Lund J-H<br />

(Eds), Proceed<strong>in</strong>gs from <strong>the</strong> CONNECT workshop ‘Physical<br />

habitat restoration <strong>in</strong> canalised watercourses – possibilities<br />

and constra<strong>in</strong>ts, Norweigan Institute of Nature Research, Oslo,<br />

Norway, 91–99.<br />

Gevrey M, Dimopoulosb I, Leka S. 2003. Modell<strong>in</strong>g <strong>the</strong> structure<br />

of aquatic communities: concepts, methods and problems. Ecological<br />

Modell<strong>in</strong>g 160: 249–164.<br />

Ghanem A, Steffl er P, Hicks F, Katopodis C. 1996. Two-dimensional<br />

hydraulic simulation of physical habitat conditions <strong>in</strong><br />

fl ow<strong>in</strong>g streams. Regulated <strong>River</strong>s-Research & Management<br />

12: 185–200.<br />

Gilvear DJ, Heal KV, Stephen A. 2002. Hydrology and <strong>the</strong> ecological<br />

quality of Scottish river ecosystems. The Science of <strong>the</strong><br />

Total Environment 294: 131–159.<br />

Gregory S, Boyer K, Gurnell AM (Eds). 2003. The Ecology and<br />

Management of Wood <strong>in</strong> World <strong>River</strong>s. American Fisheries<br />

Society Symposium 37, American Fisheries Society: Be<strong>the</strong>sda,<br />

Maryland.<br />

Gries G, Juanes F. 1998. Microhabitat use by juvenile Atlantic<br />

salmon (Salmo salar) shelter<strong>in</strong>g dur<strong>in</strong>g <strong>the</strong> day <strong>in</strong> summer.<br />

Canadian Journal of Zoology 76: 1441–1449.<br />

Gore JA, Crawford DJ, Addison DS. 1998. An analysis of artifi cial<br />

riffl es and enhancement of benthic community diversity by<br />

physical habitat simulation (PHABSIM) and direct observation.<br />

Regulated <strong>River</strong>s: Research and Management 14: 69–77.<br />

Graf WL. 2001. Damage control: restor<strong>in</strong>g <strong>the</strong> physical <strong>in</strong>tegrity<br />

of America’s rivers. Annals of <strong>the</strong> Association of American<br />

Geographers 91: 1–27.<br />

Hardy RJ, Lane SN, Ferguson RI, Parsons DR. 2003. Assess<strong>in</strong>g<br />

<strong>the</strong> credibility of a series of computational fl uid dynamic simulationsof<br />

open channel fl ow. Hydrological Processes 17:<br />

1539–60.<br />

Hardy TB. 1998. The future of habitat model<strong>in</strong>g and <strong>in</strong>stream<br />

fl ow assessment techniques Regulated <strong>River</strong>s: Research and<br />

Management 14: 405–420.<br />

Hayes JW, Stark JD, Shearer KA. 2000. Development and test of<br />

a whole-lifetime forag<strong>in</strong>g and bioenergetics growth model for<br />

drift feed<strong>in</strong>g brown trout. Transactions of <strong>the</strong> American Fisheries<br />

Society 125: 315–332.<br />

Hearne J, Johnson IW, Armitage PD. 1994. Determ<strong>in</strong>ation of<br />

ecologically acceptable fl ows <strong>in</strong> rivers with seasonal changes<br />

<strong>in</strong> <strong>the</strong> density of macrophyte Regulated <strong>River</strong>s: Research and<br />

Management 9: 117–184.<br />

Hendry K, Cragg-H<strong>in</strong>e D, O’Grady MO et al. 2003. Management<br />

of habitat for rehabilitation and enhancement of salmonid<br />

stocks. Fisheries Research 12: 171–192.<br />

Hill MT, Platt S, Beschta RL. 1991. Ecological and geomorphological<br />

concepts for <strong>in</strong>stream and out of channel fl ow requirements.<br />

<strong>River</strong>s 2: 198–210.<br />

Janauer GA. 2000. Ecohydrology: fus<strong>in</strong>g concepts and scales.<br />

Ecological Eng<strong>in</strong>eer<strong>in</strong>g 16: 9–16.<br />

Johnson IW, Elliott CRN, Gustard A et al. 1993. Ecologically<br />

Acceptable Flows National <strong>River</strong>s Authority R&D Project<br />

Record 282/1/Wx: Bristol, UK.<br />

Johnson IW, Elliott CRN, Gustard A. 1995. Modell<strong>in</strong>g <strong>the</strong> effect<br />

of groundwater abstraction on salmonid habitat availability <strong>in</strong><br />

<strong>the</strong> <strong>River</strong> Allen, Dorset, England. Regulated <strong>River</strong>s: Research<br />

& Management 10: 229–238.<br />

Junk WJ, Bayley PB, Sparks RE. 1989. The fl ood pulse concept<br />

<strong>in</strong> river-fl oodpla<strong>in</strong> systems. In: Dodge DP (Ed), Proceed<strong>in</strong>gs<br />

of <strong>the</strong> International Large <strong>River</strong> Symposium (LARS), Canadian<br />

Special Publication of Fisheries and Aquatic Sciences 106,<br />

National Research Council of Canada, Ottawa; 110–127.<br />

Kemp JL, Harper DM, Crosa GA. 1999. Use of ‘functional habitats’<br />

to l<strong>in</strong>k ecology with morphology and hydrology <strong>in</strong> river<br />

rehabilitation. Aquatic Conservation: Mar<strong>in</strong>e and Freshwater<br />

Ecosystems 9: 159–178.<br />

Kemp JL, Harper DM, Crosa GA. 2000. The habitat-scale ecohydraulics<br />

of rivers. Ecological Eng<strong>in</strong>eer<strong>in</strong>g 16: 17–29.<br />

Kondolf GM, Larsen EW, Williams JG. 2000. Measur<strong>in</strong>g and<br />

modell<strong>in</strong>g <strong>the</strong> hydraulic environment for assess<strong>in</strong>g <strong>in</strong>stream<br />

fl ows. North American Journal of Fisheries Management 20:<br />

1016–1028.<br />

Kirkby MJ. 1993. Network hydrology and geomorphology. In:<br />

Beven K, Kirkby MJ (Eds), Channel Network Hydrology. John<br />

Wiley & Sons Ltd: Chichester, UK; 1–11.<br />

K<strong>in</strong>g JM, Tharme RE, de Villiers MS (Eds). 2000. Environmental<br />

fl ow assessments for rivers: manual for <strong>the</strong> Build<strong>in</strong>g Block<br />

Methodology. Water Research Commission Report TT 131/00:<br />

Pretoria, South Africa.


Hydrological and Hydraulic Aspects of <strong>River</strong> <strong>Restoration</strong> <strong>Uncerta<strong>in</strong>ty</strong> for Ecological Purposes 137<br />

Lamouroux N, Capra H, Pouilly M. 1998. Predict<strong>in</strong>g Habitat<br />

Suitability for lotic fi sh: l<strong>in</strong>k<strong>in</strong>g statistical hydraulic models<br />

with multivariate habitat use models. Regulated <strong>River</strong>s 14:<br />

1–11.<br />

Lamouroux N, Doutriaux E, Terrier C, Zylberblat M. 1999.<br />

Modélisation des impacts de la gestion des débits réservés du<br />

Rhône sur les peuplements piscicoles. Bullet<strong>in</strong> Français de la<br />

Pêche et de la Pisciculture 352: 45–61.<br />

Lane SN. 1998. Hydraulic modell<strong>in</strong>g <strong>in</strong> hydrology and geomorphology:<br />

a review of high resolution approaches. Hydrological<br />

Processes 12: 1131–1150.<br />

Lane SN, Richards KS. 2001. The ‘validation of hydrodynamic<br />

models: some critical perspectives. In: Anderson MG, Poates<br />

PD (Eds), Model validation: perspectives <strong>in</strong> hydrological<br />

science. John Wiley & Sons Ltd: Chichester, UK; 414–438.<br />

LeClerc 2002. Ecohydraulics, last frontier for fl uvial hydraulics:<br />

research challenges and multidiscipl<strong>in</strong>ary perspectives. In:<br />

Bousmar D, Zech Y (Eds), <strong>River</strong> Flow 2002. Swets & Zeitl<strong>in</strong>ger/<br />

Balkema: Liss; 13–26.<br />

Logan P, Furse M. 2002. Prepar<strong>in</strong>g for <strong>the</strong> European Water<br />

Framework Directive – mak<strong>in</strong>g <strong>the</strong> l<strong>in</strong>ks between habitat and<br />

aquatic biota. Aquatic Conservation: Mar<strong>in</strong>e and Freshwater<br />

Ecosystems 12: 425–437.<br />

Lytle DA, Poff NL. 2004. Adaptation to natural fl ow regimes.<br />

Trends <strong>in</strong> Ecology and Evolution 19: 94–100.<br />

Maddock I. 1999. The importance of physical habitat assessment<br />

for evaluat<strong>in</strong>g river health. Freshwater Biology 41: 373–391.<br />

Maddock IP, Bird D. 1996. The application of habitat mapp<strong>in</strong>g to<br />

identify representative PHABSIM sites on <strong>the</strong> <strong>River</strong> Tavy,<br />

Devon, UK. In: Leclerc et al. (Eds), Proceed<strong>in</strong>gs of <strong>the</strong> 2nd<br />

International Symposium on habitats and hydraulics, Quebec,<br />

Canada, June 1996.<br />

Marsh TJ. 2002. Capitalis<strong>in</strong>g on river fl ow data to meet chang<strong>in</strong>g<br />

national needs – a UK perspective. Flow Measurement and<br />

Instrumentation 13: 291–298.<br />

Mat<strong>the</strong>ws RC, Bao Y. 1991. The Texas method of prelim<strong>in</strong>ary<br />

<strong>in</strong>stream fl ow determ<strong>in</strong>ation. <strong>River</strong>s 2: 295–310.<br />

McBa<strong>in</strong> & Trush, Inc. (Eds). 2002. San Joaqu<strong>in</strong> <strong>River</strong> <strong>Restoration</strong><br />

Study Background Report. Prepared for Friant Water Users<br />

Authority, L<strong>in</strong>dsay, CA, and Natural Resources Defense<br />

Council: San Francisco, California.<br />

McPherson JI. 1997. Appeal by Thames Water Utilities under<br />

Section 43 of <strong>the</strong> Water Resources Act 1991, Inspector’s<br />

Report, WAT/95/22. Department of <strong>the</strong> Environment: Bristol,<br />

UK.<br />

Mesick CF. 1988. Effects of food and cover on numbers of apache<br />

and brown trout establish<strong>in</strong>g residency <strong>in</strong> artifi cial stream channels.<br />

Transactions of <strong>the</strong> American Fisheries Society 117:<br />

421–431.<br />

Milhous RT. 1999. History, <strong>the</strong>ory, use, and limitations of <strong>the</strong><br />

Physical Habitat Simulation System. Proceed<strong>in</strong>gs of <strong>the</strong> 3rd<br />

International Symposium on Ecohydraulics, Salt Lake City,<br />

Utah, USA. Available on CD-ROM only.<br />

Milhous RT, Updike MA, Scheider DM. 1989. Physical Habitat<br />

Simulation (PHABSIM) Reference Manual, Version II.<br />

Instream Flow Information Paper No. 26. US Fish and Wildlife<br />

Service Biology Report 89 (16).<br />

Morhardt JE, Hanson DF, Coulston CJ. 1983. Instream fl ow<br />

analysis: Increased accuracy us<strong>in</strong>g habitat mapp<strong>in</strong>g. Waterpower<br />

83- An International Conference on Hydropower.<br />

Conference Proceed<strong>in</strong>gs Vol 3, Tennessee Valley Authority,<br />

Knoxville; 1294–1304.<br />

Mousouridis A. 2001. Identify<strong>in</strong>g Upland Channel Network<br />

Dynamics Over Decadal and Century Timescales. Unpublished<br />

PhD <strong>the</strong>sis, University of London, UK.<br />

Naiman RJ, Lonzarich DG, Beechie TJ, Ralph SC. 1992. General<br />

pr<strong>in</strong>ciples of classifi cation and <strong>the</strong> assessment of conservation<br />

potential <strong>in</strong> rivers. In: Boon PJ, Calow P, Petts GE (Eds), <strong>River</strong><br />

Conservation and Management. John Wiley & Sons Ltd:<br />

Chichester, UK; 93–123.<br />

Nestler J, Sutton VK. 2000. Describ<strong>in</strong>g scales of features <strong>in</strong> river<br />

channels us<strong>in</strong>g fractal geometry concepts. Regulated <strong>River</strong>s:<br />

Research & Management 16: 1–22.<br />

Newbury R, Gaboury M. 1993. Exploration and rehabilitation of<br />

hydraulic habitats <strong>in</strong> streams us<strong>in</strong>g pr<strong>in</strong>ciples of fl uvial behaviour.<br />

Freshwater Biology 29: 195–210.<br />

Newson MD. 2002. Geomorphological concepts and tools for<br />

susta<strong>in</strong>able river ecosystem management. Aquatic Conservation:<br />

Mar<strong>in</strong>e and Freshwater Ecosystems 12: 365–379.<br />

Newson MD, Newson CL. 2000. Geomorphology, ecology<br />

and river channel habitat: mesoscale approaches to bas<strong>in</strong>scale<br />

challenges. Progress <strong>in</strong> Physical Geography 24: 195–<br />

217.<br />

Newson MD, Harper DM, Padmore CL et al. 1998. A costeffective<br />

approach for l<strong>in</strong>k<strong>in</strong>g habitats, fl ow types and species<br />

requirements. Aquatic Conservation: Mar<strong>in</strong>e and Freshwater<br />

Ecosystems 8: 431–446.<br />

Nicholas AP. 2001. Computational fl uid dynamics modell<strong>in</strong>g of<br />

boundary roughness <strong>in</strong> gravel-bed rivers: an <strong>in</strong>vestigation of<br />

<strong>the</strong> effects of random variability <strong>in</strong> bed elevation. Earth Surface<br />

Processes and Landforms 26: 345–362.<br />

Oreskes N, Shrader-Frechette K, Belitz K. 1994. Verifi cation,<br />

validation, and confi rmation of numerical models <strong>in</strong> <strong>the</strong> Earth<br />

Sciences. Science 263: 641–646.<br />

Padmore CL. 1997. Biotopes and <strong>the</strong>ir hydraulics: a method for<br />

defi n<strong>in</strong>g <strong>the</strong> physical component of freshwater quality. In:<br />

Boon PJ, Howell DL (Eds), Freshwater Quality: Defi n<strong>in</strong>g<br />

<strong>the</strong> Indefi nable? The Stationery Offi ce: Ed<strong>in</strong>burgh, UK;<br />

251–257.<br />

Parasiewicz P. 2001. MesoHABSIM: A concept for application<br />

of <strong>in</strong>stream fl ow models <strong>in</strong> river restoration plann<strong>in</strong>g. Fisheries<br />

26: 6–13.<br />

Parasiewicz P, Dunbar MJ. 2001. Physical habitat modell<strong>in</strong>g<br />

for fi sh: A develop<strong>in</strong>g approach. Large <strong>River</strong>s, 12 2–4, Arch.<br />

Hydrobiol. Suppl. 135/2–4: 239–268.<br />

Peters MR, Abt SR, Watson CC et al. 1995. Assessment of<br />

restored river<strong>in</strong>e habitat us<strong>in</strong>g RCHARC. Water Resources<br />

Bullet<strong>in</strong> 31: 745–752<br />

Petts GE. 1984. Impounded <strong>River</strong>s: Perspectives for Ecological<br />

Management. John Wiley & Sons Ltd: Chichester, UK.<br />

Petts GE, Amoros C. 1996. Fluvial Hydrosystems. Chapman and<br />

Hall: London, UK.<br />

Poff NL, Allan JD, Ba<strong>in</strong> MB et al. 1997. The natural fl ow regime.<br />

BioScience 47: 769–784.


138 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Prudhomme C, Jakob D, Svensson C. 2003. <strong>Uncerta<strong>in</strong>ty</strong> and<br />

climate change impact on <strong>the</strong> fl ood regime of small UK catchments.<br />

Journal of Hydrology 277: 1–23.<br />

Pusey BJ. 1998. Methods address<strong>in</strong>g <strong>the</strong> fl ow requirements of<br />

fi sh. In: Arth<strong>in</strong>gton AH, Zalucki JM (Eds), Comparative evaluation<br />

of environmental fl ow assessment techniques review of<br />

methods. Occassional Paper 27/98, Land and Water Resources<br />

and Development Corporation: Canberra, Australia; 64–103.<br />

Railsback SF, Lamberson RH, Harvey BC, Duffy WE, 1999.<br />

Movement rules for <strong>in</strong>dividual-based models of stream fi sh.<br />

Ecological Modell<strong>in</strong>g 123: 73–89.<br />

Reice SR, Wissmar RC, Naiman RJ. 1990. Disturbance regimes,<br />

resilience and recovery of animal communities and habitats <strong>in</strong><br />

lotic ecosystems. Environmental Management 14: 647–660.<br />

Reyjol Y, Lim P, Belaud A, Lek S. 2001. Modell<strong>in</strong>g microhabitat<br />

used by fi sh <strong>in</strong> natural and regulated fl ows <strong>in</strong> <strong>the</strong> river Garonne<br />

(France). Ecological Modell<strong>in</strong>g 146: 131–142.<br />

Reynolds CS. 2002. Ecological pattern and ecosystem <strong>the</strong>ory.<br />

Ecological Modell<strong>in</strong>g 158: 181–200.<br />

Richter BD, Baumgartner JV, Powell J, Braun DP. 1996. A method<br />

for assess<strong>in</strong>g hydrological alteration with<strong>in</strong> ecosystems. Conservation<br />

Biology 10: 1163–1174.<br />

Richter BD, Baumgartner JV, Wig<strong>in</strong>gton R, Braun DP. 1997. How<br />

much water does a river need? Freshwater Biology 37:<br />

231–249.<br />

Rodriguez-Iturbe, Valdes JB. 1979. The geomorphic structure<br />

of hydrologic response. Water Resources Research 15:<br />

1409–1420.<br />

Roussel JM, Bardonnet A, Claude A. 1999. Microhabitats of<br />

brown trout when feed<strong>in</strong>g on drift and when rest<strong>in</strong>g <strong>in</strong> a<br />

lowland salmonid brook: effects on Weighted Usable Area.<br />

Archiv Fur Hydrobiologie 146: 413–429.<br />

Rowntree K, Wadeson R. 1998. A geomorphological framework<br />

for <strong>the</strong> assessment of <strong>in</strong>stream fl ow requirements. Aquatic Ecosystem<br />

Health and Management 1: 125–141.<br />

Sabaton C, Souchon Y, Lascaux JM et al. 2003. The ‘Guaranteed<br />

Flow Work<strong>in</strong>g Group’: A French feedback of IFIM based on<br />

habitat and brown trout population time series observations. In:<br />

Lamb BL, D Garcia de Jalon C, Sabaton C et al. (Eds), Proceed<strong>in</strong>gs<br />

of International IFIM users conference: Colorado<br />

State University, Fort Coll<strong>in</strong>s, Colorado.<br />

Schmidt JC, Webb RH, Valdez RA et al. 1998. Science and values<br />

<strong>in</strong> river restoration <strong>in</strong> <strong>the</strong> Grand Canyon. BioScience 48:<br />

735–747.<br />

Smakht<strong>in</strong> VU. 2001. Low fl ow hydrology: a review. Journal of<br />

Hydrology 240: 147–186.<br />

Souchon Y, Capra H. 2003. Aquatic habitat modell<strong>in</strong>g: biological<br />

validations of IFIM/PHABSIM methodology and new perspectives<br />

In: Lamb BL, D Garcia de Jalon C, Sabaton C et al. (Eds),<br />

Proceed<strong>in</strong>gs of International IFIM users conference: Colorado<br />

State University, Fort Coll<strong>in</strong>s, Colorado.<br />

Souchon Y, Keith P. 2001. Freshwater fi sh habitat: science, management<br />

and conservation <strong>in</strong> France. Aquatic Ecosystem Health<br />

and Management 4: 401–412.<br />

Spence R, Hickley P. 2000. The use of PHABSIM <strong>in</strong> <strong>the</strong> management<br />

of water resources and fi sheries <strong>in</strong> England and Wales.<br />

Ecological Eng<strong>in</strong>eer<strong>in</strong>g 16: 153–158.<br />

Stanford JA, Ward JV, Liss WJ et al. 1996. A general protocol for<br />

restoration of regulated rivers. Regulated <strong>River</strong>s: Research and<br />

Management 12: 391–413.<br />

Stillwater Sciences 2001. Tuolumne <strong>River</strong> restoration program<br />

summary report, summary of studies, conceptual models, restoration<br />

projects, and ongo<strong>in</strong>g monitor<strong>in</strong>g. Prepared for <strong>the</strong><br />

CALFED/AFRP Adaptive Management Forum, with assistance<br />

from <strong>the</strong> Tuolumne <strong>River</strong> Technical Advisory Committee.<br />

Strevens AP. 1999. Impacts of groundwater abstraction on <strong>the</strong><br />

trout fi shery of <strong>the</strong> <strong>River</strong> Piddle; and an approach to <strong>the</strong>ir alleviation.<br />

Hydrological Processes 13: 487–496.<br />

Tennant DL. 1976. Instream fl ow regimens for fi sh, wildlife,<br />

recreation and related environmental resources. Fisheries 1:<br />

6–10.<br />

Thoms MC, Sheldon F. 2002. An ecosystem approach for determ<strong>in</strong><strong>in</strong>g<br />

environmental water allocations <strong>in</strong> Australian dryland<br />

river systems: <strong>the</strong> role of geomorphology. Geomorphology 47:<br />

153–168.<br />

Turnpenny AWH, Blay SR, Carron JJ, Clough SC. 2001. Literature<br />

Review of Swimm<strong>in</strong>g Speeds of Freshwater Fish. Environment<br />

Agency R&D Technical Report W2–026/TR2,<br />

Environment Agency: Bristol, UK.<br />

US Army Corps of Eng<strong>in</strong>eers. 2002. HEC-RAS <strong>River</strong> Analysis<br />

System. CDP-68. US Army Corps of Eng<strong>in</strong>eers: Davis,<br />

California.<br />

Valdimarsson SK, Metcalfe NB. 2001. Is <strong>the</strong> level of aggression<br />

and dispersion <strong>in</strong> territorial fi sh dependent on light <strong>in</strong>tensity?<br />

Animal Behaviour 61: 1143–1149.<br />

Van W<strong>in</strong>kle W, Jager HI, Railsback SF et al. 1998. Individualbased<br />

model of sympatric populations of brown and ra<strong>in</strong>bow<br />

trout for <strong>in</strong>stream fl ow assessment: model description and calibration.<br />

Ecological Modell<strong>in</strong>g 110: 175–207.<br />

Vismara R, Azzell<strong>in</strong>o A, Bosi R et al. 2001. Habitat suitability<br />

curves for brown trout (Salmo trutta fario L.) <strong>in</strong> <strong>the</strong> <strong>River</strong> Adda,<br />

Nor<strong>the</strong>rn Italy: Compar<strong>in</strong>g univariate and multivariate approaches.<br />

Regulated <strong>River</strong>s, Research & Management 17: 37–50.<br />

Waters BF. 1976. A methodology for evaluat<strong>in</strong>g <strong>the</strong> effects of<br />

different stream fl ows on salmonid habitat. In: Orsborn JF,<br />

Allman CH (Eds), Instream Flow Needs. American Fisheries<br />

Society: Be<strong>the</strong>sda, Maryland. 254–266.<br />

Werner H, Oback M. 2001. New neural network types estimat<strong>in</strong>g<br />

<strong>the</strong> accuracy response for ecological modell<strong>in</strong>g. Ecological<br />

Modell<strong>in</strong>g 146: 289–298.<br />

Wood PJ, Hannah DM, Agnew MD, Petts GE. 2001. Scales of<br />

hydroecological variability with<strong>in</strong> a groundwater-dom<strong>in</strong>ated<br />

stream. Regulated <strong>River</strong>s: Research and Management 17:<br />

347–367.<br />

Wu FC, Wang CF. 2002. Effect of fl ow-related substrate alteration<br />

on physical habitat: A case study of <strong>the</strong> endemic river loach<br />

S<strong>in</strong>ogastromyzon puliensis (Cypr<strong>in</strong>iformes, Homalopteridae)<br />

downstream of Chi-Chi diversion weir, Chou-Shui Creek,<br />

Taiwan. <strong>River</strong> Research and Applications 18: 155–169.<br />

Yrjänä T, Lahti M, Kamula R. 1999. Effects of a habitat enhancement<br />

experiment <strong>in</strong> <strong>the</strong> Laukka area of <strong>the</strong> river Oulujoki,<br />

F<strong>in</strong>land. University of Oulu. Publications of Water Resources<br />

and Environmental Eng<strong>in</strong>eer<strong>in</strong>g Laboratory. A5. (In F<strong>in</strong>nish,<br />

with English Summary).


<strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat Edited by Stephen Darby and David Sear<br />

© 2008 John Wiley & Sons, Ltd<br />

8<br />

<strong>Uncerta<strong>in</strong>ty</strong> Surround<strong>in</strong>g <strong>the</strong> Ecological<br />

Targets and Response of <strong>River</strong> and<br />

Stream <strong>Restoration</strong><br />

8.1 INTRODUCTION<br />

Mart<strong>in</strong> R. Perrow 1 , Eleanor R. Skeate 1 , David Leem<strong>in</strong>g 2 , Judy England 3 and<br />

Mark L. Toml<strong>in</strong>son 1<br />

In his sem<strong>in</strong>al paper, Bradshaw (1987) described restoration<br />

ecology as <strong>the</strong> ‘acid test’ of ecology <strong>in</strong> that if <strong>the</strong>re<br />

is suffi cient understand<strong>in</strong>g of form, structure, process and<br />

function <strong>the</strong>n it should be possible to restore any particular<br />

habitat to a close approximation of that which is desired.<br />

The recent Handbook of Ecological <strong>Restoration</strong> <strong>in</strong> two<br />

volumes detail<strong>in</strong>g pr<strong>in</strong>ciples (Perrow and Davy, 2002a)<br />

and practice (Perrow and Davy, 2002b) suggests that restoration<br />

ecology has come of age as a science, although<br />

<strong>the</strong>re is great disparity <strong>in</strong> <strong>the</strong> level of understand<strong>in</strong>g and<br />

thus success <strong>in</strong> different biomes. The book also reveals<br />

<strong>the</strong> fundamentally different approach to <strong>the</strong> restoration of<br />

lotic and lentic environments.<br />

In lakes, especially shallow ones, <strong>the</strong>re has been a <strong>the</strong>oretical<br />

shift from ‘bottom-up’, where <strong>the</strong> form and function<br />

was controlled by physical and chemical properties,<br />

to ‘top-down’, where components at <strong>the</strong> top of <strong>the</strong> food<br />

web, especially fi sh, have great bear<strong>in</strong>g on lower trophic<br />

levels and even physical properties (Perrow et al., 2002).<br />

The focus of lake restoration is <strong>the</strong>refore often on fi sh<br />

(Jeppesen and Sammalkarpi, 2002) embrac<strong>in</strong>g and us<strong>in</strong>g<br />

fundamentally ecological <strong>in</strong>teractions.<br />

In rivers and streams <strong>the</strong>re has also been a shift <strong>in</strong> focus<br />

away from water quality – which is now not seen as <strong>the</strong><br />

primary limit<strong>in</strong>g factor for a healthy, natural system as it<br />

was immediately after <strong>the</strong> <strong>in</strong>dustrial revolution and prior<br />

1 ECON, Ecological Consultancy, UK.<br />

2 Consultant Ecologist, UK<br />

3 Environment Agency, UK<br />

to <strong>the</strong> development of improved water quality standards –<br />

to <strong>the</strong> limitations of physical habitat structure. For example<br />

<strong>in</strong> <strong>the</strong> United K<strong>in</strong>gdom, <strong>in</strong>itial restoration efforts on <strong>the</strong><br />

<strong>River</strong> Thames <strong>in</strong> 1858 were driven by water problems<br />

caused by pollution from human and animal wastes<br />

(Gameson and Wheeler, 1977), whereas now <strong>the</strong> focus is<br />

on restor<strong>in</strong>g habitat quality, especially for fi sh, <strong>in</strong> <strong>the</strong> upper<br />

catchment where tributaries have suffered from river eng<strong>in</strong>eer<strong>in</strong>g<br />

schemes (Rob<strong>in</strong>son and Whitton, 2004).<br />

What has become <strong>the</strong> billion dollar <strong>in</strong>dustry of river<br />

restoration (Malakoff, 2004; Palmer et al., 2005) generally<br />

uses a geomorphological (<strong>in</strong> comb<strong>in</strong>ation with eng<strong>in</strong>eer<strong>in</strong>g)<br />

approach (Downs et al., 2002; Malakoff 2004), <strong>in</strong> <strong>the</strong><br />

hierarchical gradient of process–form–habitat–biota promoted<br />

by <strong>the</strong> highly <strong>in</strong>fl uential National Research Council<br />

(of <strong>the</strong> United States) publication <strong>Restoration</strong> of Aquatic<br />

Ecosystems (1992). This is restore natural water and sediment<br />

regime; restore natural channel geometry; restore<br />

natural riparian plant communities; and restore native<br />

aquatic plants and animals.<br />

The widely perceived ecological focus of river restoration<br />

(Newson et al., 2002; Downs and Sk<strong>in</strong>ner 2002;<br />

Ormerod, 2004), is <strong>the</strong>refore <strong>the</strong> fi nal step <strong>in</strong> <strong>the</strong> cha<strong>in</strong> of<br />

works. As with ripples on a pond or <strong>the</strong> cascade response<br />

through trophic levels (Carpenter and Kitchell, 1993), <strong>the</strong><br />

strength of <strong>the</strong> <strong>in</strong>teraction and <strong>the</strong> response is likely to<br />

weaken <strong>the</strong> fur<strong>the</strong>r from <strong>the</strong> source action. To take a more<br />

specifi c example, <strong>the</strong> huge amount of river restoration


140 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

focused on anadromous fi shes <strong>in</strong>clud<strong>in</strong>g sturgeon<br />

(Waldman and Wirg<strong>in</strong>, 1998) and particularly salmonids,<br />

upon which millions of dollars are spent annually <strong>in</strong> <strong>the</strong><br />

US Pacifi c Northwest alone (Roni et al., 2002), effectively<br />

h<strong>in</strong>ges on <strong>the</strong> relationship between <strong>the</strong> fi sh and limit<strong>in</strong>g<br />

habitat variables. However, such relationships are not<br />

always clear-cut and may expla<strong>in</strong> a low proportion of <strong>the</strong><br />

variation <strong>in</strong> abundance and biomass (Milner et al., 1985).<br />

In very simple terms, a ‘restored’ river may be perfectly<br />

capable of support<strong>in</strong>g a high abundance and biomass of<br />

fi sh but may actually conta<strong>in</strong> very few. In recognition of<br />

<strong>the</strong> fact that biological and ecological limitations (i.e.<br />

stock limited recruitment, disease, predation, competition<br />

etc) may be more important than <strong>the</strong> generic problem of<br />

not hav<strong>in</strong>g enough habitat, release of large numbers of<br />

artifi cially raised juveniles may have to be undertaken, e.g.<br />

on <strong>the</strong> <strong>River</strong> Mattole <strong>in</strong> California (Mattole <strong>Restoration</strong><br />

council http://www.mattole.org/) and <strong>in</strong> rivers <strong>in</strong> North-<br />

East England (Russell, 1994). In a nutshell, <strong>the</strong> paradigm<br />

of gett<strong>in</strong>g <strong>the</strong> physical structure right and <strong>the</strong>n <strong>the</strong> ‘ecology’<br />

will <strong>the</strong>n surely follow is likely to be a false premise<br />

(Ormerod, 2004).<br />

Geomorphological and ecological processes are <strong>in</strong>extricably<br />

<strong>in</strong>tertw<strong>in</strong>ed and best considered <strong>in</strong> an eco-hydromorphic<br />

approach to restoration (Clarke et al., 2003). To<br />

illustrate, consider a low order boulder-strewn upland<br />

stream. Here, geological structure, gradient and geomorphological<br />

processes expressed <strong>in</strong> fl ow and sediment<br />

regime may <strong>in</strong>itially be viewed as dom<strong>in</strong>ant over ecological<br />

processes. But consider <strong>the</strong> impact of <strong>the</strong> presence or<br />

absence of trees and <strong>the</strong>ir coarse woody debris (CWD)<br />

feed<strong>in</strong>g back <strong>in</strong>to geomorphological processes by constra<strong>in</strong><strong>in</strong>g<br />

channel width and <strong>in</strong>creas<strong>in</strong>g fl ows and promot<strong>in</strong>g<br />

erosion <strong>in</strong> one place and slow<strong>in</strong>g fl ows and caus<strong>in</strong>g<br />

deposition <strong>in</strong> ano<strong>the</strong>r. Allochthonous <strong>in</strong>put <strong>in</strong> <strong>the</strong> form of<br />

leaf fall provides coarse particulate matter (CPOM) to <strong>the</strong><br />

channel provid<strong>in</strong>g <strong>the</strong> basis of nutrient cycl<strong>in</strong>g and spiral<strong>in</strong>g<br />

(see Newbold, 1992 for an overview) and energy fl ow<br />

(Calow, 1992 for an overview) and thus biological production,<br />

ultimately determ<strong>in</strong>es <strong>the</strong> biomass of <strong>the</strong> biota.<br />

Coarse woody debris from <strong>the</strong> trees promotes retentiveness<br />

and decomposition of litter (Lepori et al., 2005), key ecological<br />

functions <strong>in</strong>fl uenc<strong>in</strong>g <strong>the</strong> abundance and diversity<br />

of shredd<strong>in</strong>g <strong>in</strong>vertebrates responsible for produc<strong>in</strong>g fi ne<br />

particulate matter (FPOM) used by o<strong>the</strong>r feed<strong>in</strong>g guilds,<br />

consumed <strong>in</strong> turn by fi sh, birds and mammals (Richardson<br />

and Jackson, 2002). In this hypo<strong>the</strong>tical example, understand<strong>in</strong>g<br />

and subsequently manipulat<strong>in</strong>g ecological processes,<br />

such as succession, herbivory (predation) and<br />

competition, to <strong>in</strong>fl uence <strong>the</strong> structure and composition of<br />

<strong>the</strong> riparian tree fl ora may be as (if not more) effective as<br />

manipulat<strong>in</strong>g physical habitat structure.<br />

Moreover, whilst ecological processes may be embodied<br />

with<strong>in</strong> a particular component of <strong>the</strong> fauna and fl ora, this<br />

is <strong>in</strong>tuitively unlikely to operate over trophic scales with<strong>in</strong><br />

biotic components as well as <strong>in</strong>teract with abiotic components.<br />

Individuals, species populations and <strong>the</strong>ir communities<br />

cannot be seen as mere products of <strong>the</strong> physical (and<br />

chemical) framework but <strong>in</strong>extricably l<strong>in</strong>ked with it via<br />

particular processes. Thus, an eng<strong>in</strong>eer<strong>in</strong>g analogue of<br />

fi tt<strong>in</strong>g habitat components back toge<strong>the</strong>r (i.e. a build<strong>in</strong>g<br />

block functional habitat approach; Harper et al., 1992) to<br />

make a naturally function<strong>in</strong>g ecosystem is likely to be too<br />

simplistic. The nub of an eng<strong>in</strong>eer<strong>in</strong>g approach, that a<br />

habitat component or unit may be expressed as a certa<strong>in</strong><br />

number of <strong>in</strong>dividuals or composition of a community,<br />

thus takes no account of process, but ra<strong>the</strong>r assumes it.<br />

Our experiences of river restoration <strong>in</strong> <strong>the</strong> United<br />

K<strong>in</strong>gdom lead us to believe <strong>the</strong>re is a widespread lack of<br />

understand<strong>in</strong>g of what ‘ecology’ is and does <strong>in</strong> river restoration.<br />

Pert<strong>in</strong>ent to this volume, we aim to explore<br />

whe<strong>the</strong>r this <strong>in</strong>tuitively leads to uncerta<strong>in</strong>ty (<strong>in</strong> this case<br />

reducible ignorance, see Chapter 3) and a lack of confi -<br />

dence <strong>in</strong> ecologists amongst o<strong>the</strong>r types of restoration<br />

practitioner. For example, Downs and Sk<strong>in</strong>ner (2002) suggested<br />

‘river restoration ecology suffers from <strong>the</strong> <strong>in</strong>ability<br />

of ecologists to defi ne <strong>the</strong>ir system requirements as closely<br />

as, for <strong>in</strong>stance, eng<strong>in</strong>eers can specify and model (although<br />

not necessarily achieve) <strong>the</strong>ir fl ood defence requirements’.<br />

This neatly encapsulates <strong>the</strong> perception that not only is<br />

ecological restoration uncerta<strong>in</strong> but that <strong>the</strong> scientifi c basis<br />

to predict and subsequently evaluate ecological response<br />

is severely lack<strong>in</strong>g, perhaps even absent.<br />

One of <strong>the</strong> aims of this chapter is to determ<strong>in</strong>e whe<strong>the</strong>r<br />

such bold statements are justifi ed <strong>in</strong> <strong>the</strong> context of <strong>the</strong><br />

basic premise of this volume that much restoration of<br />

rivers has and is be<strong>in</strong>g conducted with considerable uncerta<strong>in</strong>ty<br />

<strong>in</strong> one form or ano<strong>the</strong>r. To do this, <strong>the</strong> sources of<br />

uncerta<strong>in</strong>ty surround<strong>in</strong>g ecology (and ecologists) affect<strong>in</strong>g<br />

river restoration schemes are explored <strong>in</strong> stepwise manner,<br />

broadly ak<strong>in</strong> to <strong>the</strong> plann<strong>in</strong>g, design and target sett<strong>in</strong>g,<br />

implementation and outcome and appraisal phases <strong>in</strong> any<br />

project (Figure 8.1). An attempt is made to determ<strong>in</strong>e if<br />

<strong>the</strong>re is widespread uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> ecological response<br />

and outcome of restoration schemes over and above that<br />

expected as a result of natural variation, and, if so, whe<strong>the</strong>r<br />

uncerta<strong>in</strong>ty and unreliability can be reduced.<br />

8.2 SOURCES OF ECOLOGICAL UNCERTAINTY<br />

8.2.1 Inherent Variability<br />

Natural systems are <strong>in</strong>herently variable <strong>in</strong> time and space,<br />

and probably none more so than rivers. Variation <strong>in</strong> space


Figure 8.1 Is Ecological <strong>River</strong> <strong>Restoration</strong> ‘Uncerta<strong>in</strong>’? Infl uences of Institutional and Ecological Drivers <strong>in</strong> Project Plann<strong>in</strong>g and Implementation<br />

<strong>Uncerta<strong>in</strong>ty</strong> Surround<strong>in</strong>g <strong>the</strong> Ecological Targets and Response of <strong>River</strong> and Stream <strong>Restoration</strong> 141


142 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

is particularly well developed both longitud<strong>in</strong>ally, laterally<br />

and vertically <strong>in</strong> rivers, as recognised by attempts to<br />

understand <strong>the</strong>m with<strong>in</strong> <strong>the</strong> river cont<strong>in</strong>uum concept<br />

(RCC) (Vannote et al., 1980). In larger rivers with extensive<br />

fl oodpla<strong>in</strong> systems <strong>the</strong> fl ood pulse concept (Junk et<br />

al., 1989) has been used to frame <strong>in</strong>teractions between<br />

fl oodpla<strong>in</strong> and channel, which typically vary over time.<br />

Temporal variation may also conta<strong>in</strong> o<strong>the</strong>r important <strong>in</strong>terannual<br />

(random, periodic or semi-periodic), annual, seasonal<br />

and diel components, not simply amongst fauna and<br />

fl ora but also <strong>in</strong> <strong>the</strong> rates of ecosystem processes such as<br />

productivity and nutrient and energy fl ow.<br />

Buijse et al. (2002) outl<strong>in</strong>e that whilst considerable<br />

effort and emphasis have been directed at attempt<strong>in</strong>g to<br />

quantify and understand biological, hydrological and geomorphological<br />

<strong>in</strong>teractions that toge<strong>the</strong>r determ<strong>in</strong>e <strong>the</strong><br />

function<strong>in</strong>g of a river fl oodpla<strong>in</strong> system, <strong>the</strong>se are <strong>in</strong>evitably<br />

variable and complex. Complexity occurs at <strong>the</strong><br />

physical level (variation <strong>in</strong> types of river<strong>in</strong>e system), <strong>the</strong><br />

restoration level (<strong>the</strong> comb<strong>in</strong>ation of features requir<strong>in</strong>g<br />

restoration are site-specifi c) and <strong>the</strong> ecological level (community<br />

<strong>in</strong>teractions). Hughes et al. (Chapter 6) argue that<br />

‘complexity’ accounts for a degree of <strong>the</strong> perceived ecological<br />

‘uncerta<strong>in</strong>ty’ <strong>in</strong> river restoration and it must be<br />

accepted that much may rema<strong>in</strong> unknowable (Chapter 3).<br />

In a wider discussion of management of nature <strong>in</strong> <strong>the</strong><br />

United K<strong>in</strong>gdom, Adams (1997) suggested a recent shift<br />

<strong>in</strong> ecological th<strong>in</strong>k<strong>in</strong>g from look<strong>in</strong>g to restore a ‘stable<br />

equilibrium’, often taken to represent historical conditions,<br />

to a new view that a river<strong>in</strong>e system is by its nature<br />

dynamic and chang<strong>in</strong>g, i.e. one that embraces natural variability.<br />

This is <strong>the</strong> basis of ‘recovery enhancement’ where<br />

<strong>the</strong> scope for natural recovery is exploited, after <strong>the</strong> shackl<strong>in</strong>g<br />

constra<strong>in</strong>ts such as structures (e.g. weirs) and modifi<br />

ed eng<strong>in</strong>eered banks, management actions such as<br />

<strong>in</strong>tensive vegetation management regimes or even simply<br />

graz<strong>in</strong>g by livestock are removed. The river may also be<br />

given <strong>the</strong> space to move (i.e. dur<strong>in</strong>g <strong>the</strong> re-connection<br />

of <strong>the</strong> fl oodpla<strong>in</strong> as on <strong>the</strong> <strong>River</strong>s Cole and Brede, see<br />

Kronvang et al., 1998; Holmes and Nielsen, 1998). Such<br />

action to favour a potentially less controllable and predictable<br />

system (Hughes et al., Chapter 6) may appear to <strong>the</strong><br />

eng<strong>in</strong>eer, whose target is to achieve control over <strong>the</strong> system<br />

to meet specifi ed fl ood defence and shipp<strong>in</strong>g access<br />

targets, as simple acceptance of uncerta<strong>in</strong>ty.<br />

8.2.2 Limited Understand<strong>in</strong>g<br />

In spite of <strong>the</strong> wealth of work carried out on river<strong>in</strong>e<br />

systems, several authors (de Waal et al., 1995; Buijse<br />

et al., 2002) have argued that <strong>the</strong>re is actually a lack<br />

of scientifi c literature about river restoration, and that a<br />

lack of robust datasets has rendered much of <strong>the</strong> work<br />

that has been carried out unpublishable <strong>in</strong> peer reviewed<br />

journals. Although l<strong>in</strong>ked, <strong>the</strong>re are actually two issues<br />

here, <strong>in</strong>formation exchange and scientifi c substance and<br />

credibility.<br />

Information exchange has undoubtedly <strong>in</strong>creased enormously<br />

as <strong>the</strong> number of river restoration projects across<br />

<strong>the</strong> globe has <strong>in</strong>creased. Malakoff (2004) reports 30 000<br />

projects <strong>in</strong> <strong>the</strong> United States alone with tens of thousands<br />

more to come <strong>in</strong> <strong>the</strong> next few years. The literature search<br />

for this chapter uncovered a huge variety of <strong>in</strong>formation<br />

on various river restoration projects, especially via websites<br />

such as <strong>the</strong> European Centre for <strong>River</strong> <strong>Restoration</strong><br />

(http://www.rws.nl/rws/riza/home/ecrr/) and its UK<br />

(http://www.<strong>the</strong>rrc.co.uk/) and Danish counterparts (http://<br />

www2.dmu.dk/) and <strong>the</strong> National <strong>River</strong> <strong>Restoration</strong><br />

Science Syn<strong>the</strong>sis, which operates <strong>in</strong> seven states of <strong>the</strong><br />

United States, with a satellite branch <strong>in</strong> Victoria, Australia<br />

(http://www.nrrss.umd.edu/). However, much <strong>in</strong>formation<br />

is produced <strong>in</strong> an abbreviated form and it diffi cult to<br />

determ<strong>in</strong>e whe<strong>the</strong>r many projects outl<strong>in</strong>ed <strong>in</strong> such a<br />

manner are of real scientifi c substance.<br />

At <strong>the</strong> opposite end of <strong>the</strong> spectrum lies <strong>the</strong> academic<br />

peer-reviewed process as a vehicle for <strong>in</strong>formation<br />

exchange. The fact that Ormerod (2004) found only 300+<br />

papers <strong>in</strong>clud<strong>in</strong>g <strong>the</strong> terms river or stream restoration or<br />

rehabilitation <strong>in</strong> <strong>the</strong> title, abstract or key words <strong>in</strong> <strong>the</strong> ISI®<br />

database, <strong>in</strong>dicates that only a fraction of projects reach<br />

<strong>the</strong> academic literature. This could illustrate that much<br />

work is not submitted for publication and/or many submissions<br />

are not accepted as a result of <strong>in</strong>adequate scientifi c<br />

rigour. It may also suggest that few academic river ecologists<br />

are <strong>in</strong>volved with restoration. In <strong>the</strong> United K<strong>in</strong>gdom<br />

at least, ecological <strong>in</strong>put <strong>in</strong> many schemes is undertaken<br />

<strong>in</strong>ternally by statutory bodies such as <strong>the</strong> Environment<br />

Agency or by consult<strong>in</strong>g fi rms (like <strong>the</strong> authors of this<br />

chapter!).<br />

Overall, whilst <strong>the</strong>re is a proliferation of <strong>in</strong>formation,<br />

facilitat<strong>in</strong>g <strong>the</strong> rapid development of project plann<strong>in</strong>g,<br />

target sett<strong>in</strong>g, implementation and post-project appraisal<br />

on a global scale, <strong>the</strong> question-mark over data quality and<br />

<strong>the</strong> lack of <strong>in</strong>volvement of academic ecologists may mean<br />

<strong>the</strong> uncerta<strong>in</strong>ty due to limited knowledge is not reduc<strong>in</strong>g<br />

as rapidly as it might.<br />

8.2.3 Project Selection<br />

In <strong>the</strong> United K<strong>in</strong>gdom, until very recently, virtually all<br />

restoration was opportunistic despite <strong>the</strong> need for a strategic<br />

approach be<strong>in</strong>g recognised over a decade ago (ECON,<br />

1993). One of <strong>the</strong> reasons for this was simply that restoration<br />

was most effectively directed at rivers (reaches) of


<strong>Uncerta<strong>in</strong>ty</strong> Surround<strong>in</strong>g <strong>the</strong> Ecological Targets and Response of <strong>River</strong> and Stream <strong>Restoration</strong> 143<br />

highest conservation value, i.e. those closest to <strong>the</strong> predisturbance<br />

state. Investment <strong>in</strong> <strong>the</strong> most degraded<br />

reaches/rivers was misplaced and wasted resources, as<br />

<strong>the</strong>y could not achieve high status as a result of overwhelm<strong>in</strong>g<br />

and complex limit<strong>in</strong>g factors (Kern, 1992).<br />

Moreover, as shown for lakes (Benndorf, 1992), <strong>the</strong><br />

pathway from highly degraded to high quality is long and<br />

littered with <strong>the</strong> opportunity for unpredictable and <strong>in</strong>direct<br />

effects, which may constra<strong>in</strong> or even reduce ecological<br />

value (i.e. colonisation by alien species). In simple terms,<br />

long restoration trajectories (see Section 8.2.6) are more<br />

diffi cult to plan and pull off.<br />

8.2.4 Project Structure<br />

It is generally acknowledged that river restoration projects<br />

require an <strong>in</strong>tegrated and <strong>in</strong>terdiscipl<strong>in</strong>ary approach,<br />

which requires a number of experts from different fi elds<br />

work<strong>in</strong>g alongside each o<strong>the</strong>r (NRC, 1992). The scale of<br />

<strong>the</strong> project appears to have great <strong>in</strong>fl uence as whilst this<br />

may have been achieved <strong>in</strong> <strong>the</strong> case of large, high-profi le<br />

schemes (e.g. <strong>the</strong> Kissimme <strong>River</strong> <strong>Restoration</strong> Project <strong>in</strong><br />

<strong>the</strong> USA – http://www.sfwmd.gov/org/erd/krr/ – and <strong>the</strong><br />

schemes tackled by <strong>the</strong> <strong>River</strong> <strong>Restoration</strong> Centre <strong>in</strong> <strong>the</strong><br />

UK – http://www.<strong>the</strong>rrc.co.uk/) it does not appear to be<br />

rout<strong>in</strong>e. Tak<strong>in</strong>g <strong>the</strong> schemes evaluated <strong>in</strong> <strong>the</strong> UK <strong>River</strong>s<br />

and Wildlife Handbook as a sample, out of <strong>the</strong> 25 schemes<br />

where <strong>the</strong> project team was described, 80% did not employ<br />

an <strong>in</strong>terdiscipl<strong>in</strong>ary approach (RSPB et al., 1995).<br />

Resources are also clearly an issue, with a common<br />

desire to ensure that <strong>the</strong> bulk of resources available are<br />

channeled <strong>in</strong>to ‘do<strong>in</strong>g’ restoration works ra<strong>the</strong>r than be<br />

absorbed <strong>in</strong>to project <strong>in</strong>frastructure and <strong>in</strong>stitutions<br />

(Bruce-Burgess and Sk<strong>in</strong>ner, 2002). The <strong>in</strong>volvement of<br />

too many organisations and stakeholders with confl ict<strong>in</strong>g<br />

<strong>in</strong>terests may also ultimately constra<strong>in</strong> <strong>the</strong> options for<br />

restoration and <strong>the</strong> ability to undertake it (McDonald<br />

et al., 2004).<br />

From an ecological perspective it seems obvious to<br />

suggest that <strong>the</strong> more central <strong>the</strong> position of ecologists <strong>in</strong><br />

<strong>the</strong> project, <strong>the</strong> greater <strong>the</strong> chance of uncerta<strong>in</strong>ties surround<strong>in</strong>g<br />

ecological issues be<strong>in</strong>g identifi ed early <strong>in</strong> <strong>the</strong> life<br />

of <strong>the</strong> project and subsequently tackled. The perception<br />

that much river restoration has been undertaken with <strong>the</strong><br />

idea of improv<strong>in</strong>g at least some elements of <strong>the</strong> biota,<br />

especially fi sh (Bruce-Burgess and Sk<strong>in</strong>ner, 2002;<br />

Ormerod, 2004) gives <strong>the</strong> impression that such schemes<br />

are ‘front-loaded’ by ecologists. Our experience is more<br />

that ecologists enter <strong>the</strong> restoration fray to ‘count bugs’ <strong>in</strong><br />

<strong>the</strong> process–form–habitat–biota sequence (see previously),<br />

particularly where an opportunity for restoration such as<br />

on <strong>the</strong> back of a ma<strong>in</strong>tenance scheme has been taken.<br />

Clearly, it is too late to set mean<strong>in</strong>gful ecological targets<br />

if <strong>the</strong> ecologists are simply used <strong>in</strong> project monitor<strong>in</strong>g.<br />

8.2.5 Goals and Target Sett<strong>in</strong>g<br />

The nature of <strong>the</strong> goal underp<strong>in</strong>s <strong>the</strong> ecological restoration<br />

process. <strong>Restoration</strong> <strong>in</strong> its strictest sense <strong>in</strong>volves <strong>the</strong><br />

return of <strong>the</strong> structure and function of an ecosystem to a<br />

condition that existed prior to disturbance (NRC, 1992)<br />

and this forms a ready-made goal for <strong>the</strong> project (White<br />

and Walker, 1997). However, reliable historic <strong>in</strong>formation<br />

is often absent and <strong>the</strong>re is also <strong>the</strong> issue that climatic<br />

conditions, amongst o<strong>the</strong>r planetary processes, may have<br />

changed suffi ciently to mean that <strong>in</strong> ecological terms <strong>the</strong><br />

river may no longer function <strong>in</strong> <strong>the</strong> same manner, even if<br />

it had rema<strong>in</strong>ed undisturbed. The use of ecologically<br />

similar but undisturbed contemporary reference sites or<br />

even professional expert op<strong>in</strong>ion based on empirical and/<br />

or computational models, represent an alternative means<br />

of establish<strong>in</strong>g a suitable goal (White and Walker, 1997;<br />

Anderson and Dugger, 1998), although any method still<br />

requires understand<strong>in</strong>g of <strong>the</strong> current nature of <strong>the</strong> system,<br />

range of natural variation, mechanisms whereby impacts<br />

have occurred and <strong>the</strong> specifi c effects of such impacts.<br />

O<strong>the</strong>rwise, restoration may be doomed to failure.<br />

The identifi cation of causes, <strong>in</strong> particular, can be diffi -<br />

cult if <strong>the</strong>y are subtle and far removed <strong>in</strong> space and time<br />

from ecological damage (NRC, 1992). Some of <strong>the</strong> most<br />

commonly overlooked factors are: <strong>the</strong> number of stresses<br />

on biotic components; <strong>the</strong> multi-causal nature of degradation;<br />

and <strong>the</strong> diversity of result<strong>in</strong>g problems. These issues<br />

may only be resolved by adequate basel<strong>in</strong>e monitor<strong>in</strong>g.<br />

Poor basel<strong>in</strong>e data potentially leads to poor target sett<strong>in</strong>g.<br />

Sett<strong>in</strong>g and achiev<strong>in</strong>g targets ultimately provides <strong>the</strong><br />

bench-measure for <strong>the</strong> success of <strong>the</strong> project. Without a<br />

clear and appropriate goal and precise and accurate targets,<br />

<strong>the</strong>re may be a lack of benefi cial impacts, and even detrimental<br />

ones. Kondolf (1998) describes <strong>the</strong> project at Rush<br />

Creek, California, USA, as an example of this type of<br />

mistake. The aim to protect <strong>the</strong> banks from erosion was<br />

<strong>in</strong>appropriate and <strong>in</strong>consistent with geomorphological and<br />

ecological pressures at <strong>the</strong> site, and <strong>the</strong> aim to create a wet<br />

meadow to stop bank erosion was probably unrealistic due<br />

to hydrologic changes. By stopp<strong>in</strong>g bank erosion, <strong>the</strong> end<br />

result of <strong>the</strong> project was that <strong>the</strong> stream’s natural processes<br />

of recovery and re-establishment of woody riparian vegetation<br />

was arrested.<br />

These sort of experiences may have lead Downs and<br />

Sk<strong>in</strong>ner (2002) to suggest that one of <strong>the</strong> two key aspects<br />

limit<strong>in</strong>g restoration was <strong>the</strong> <strong>in</strong>ability of ecologists to set<br />

targets (<strong>the</strong> o<strong>the</strong>r was monitor<strong>in</strong>g and evaluation, see<br />

below). Certa<strong>in</strong>ly, <strong>the</strong>re does seem to be a general lack of


144 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

ecological target sett<strong>in</strong>g <strong>in</strong> projects. In <strong>the</strong> United<br />

K<strong>in</strong>gdom, an evaluation of <strong>the</strong> 40 different case studies by<br />

Holmes (1998) and a fur<strong>the</strong>r 40 <strong>in</strong> <strong>the</strong> New <strong>River</strong>s &<br />

Wildlife Handbook (RSPB et al., 1995) showed that<br />

although projects were driven by ecological goals, only<br />

around 65% actually set any sort or target, and only on<br />

one scheme were quantitative targets (based around <strong>the</strong><br />

restoration of spawn<strong>in</strong>g habitat for salmonid fi shes) set<br />

and publicised (Table 8.1). Even <strong>in</strong> <strong>the</strong> recent Cole and<br />

Skerne projects, perhaps represent<strong>in</strong>g <strong>the</strong> p<strong>in</strong>nacle of restoration<br />

<strong>in</strong> <strong>the</strong> United K<strong>in</strong>gdom to date, <strong>the</strong>re is little sign<br />

of ecological target sett<strong>in</strong>g, nei<strong>the</strong>r qualitative nor qualtitative,<br />

<strong>the</strong> benefi ts brought about by <strong>the</strong> scheme be<strong>in</strong>g<br />

assessed purely on <strong>the</strong> results of monitor<strong>in</strong>g (Holmes and<br />

Nielsen, 1998; Kronvang et al., 1998; Hoffmann et al.,<br />

1998; Biggs et al., 1998; Vivash et al., 1998).<br />

This is <strong>in</strong> sharp contrast with large scale schemes such<br />

as <strong>the</strong> Kissimmee <strong>River</strong> <strong>Restoration</strong> Project RRP (KRRP)<br />

<strong>in</strong> <strong>the</strong> USA (Trexler, 1995; Toth, 1996; Toth and Anderson,<br />

1998; Toth et al., 1998) and <strong>the</strong> Rh<strong>in</strong>e <strong>in</strong> Cont<strong>in</strong>ental<br />

Europe (Buijse et al., 2002). In <strong>the</strong> latter, ecological conditions<br />

<strong>in</strong> a similarly function<strong>in</strong>g section of <strong>the</strong> Danube were<br />

used to establish target conditions. In <strong>the</strong> former, around<br />

60 target aims were selected (performance measures), all<br />

of which had explicit goals associated with restor<strong>in</strong>g <strong>the</strong><br />

full range of structural and functional processes, and<br />

which were developed with peer review by an <strong>in</strong>dependent<br />

scientifi c review panel. Targets covered <strong>the</strong> range of ecosystem<br />

components and trophic levels and <strong>in</strong>cluded habitat<br />

characteristics (12 – hydrology, geomorphology, water<br />

quality), wetland vegetation (10), food base (13 – phytoplankton,<br />

periphyton, <strong>in</strong>vertebrates, herpetofauna) and<br />

fi sh and wildlife (25) (Whalen et al., 2002). The feasibility<br />

Table 8.1 Proportion (%) of <strong>the</strong> 80 projects reviewed by<br />

Holmes (1998) and presented <strong>in</strong> <strong>the</strong> New <strong>River</strong>s and Wildlife<br />

Handbook (1995) <strong>in</strong> England, Wales and Nor<strong>the</strong>rn Ireland<br />

satisfy<strong>in</strong>g selected criteria<br />

Criteria Holmes (1998) NRWH (1995)<br />

Wholly/partially<br />

ecologically driven<br />

65 55<br />

Sites strategically selected<br />

us<strong>in</strong>g ecological<br />

restoration criteria<br />

43 30<br />

Use of an <strong>in</strong>terdiscipl<strong>in</strong>ary<br />

team<br />

50 13<br />

With specifi c ecological<br />

targets<br />

65 68<br />

With some sort of appraisal 70 100<br />

Achiev<strong>in</strong>g some measure of<br />

ecological improvement<br />

100 92<br />

of <strong>the</strong> targets was considered <strong>in</strong> great detail. To evaluate<br />

<strong>the</strong> restoration of <strong>the</strong> fi sh community a conceptual model<br />

outl<strong>in</strong><strong>in</strong>g aspects of ecosystem function was developed<br />

(Trexler, 1995). This <strong>in</strong>tegrated <strong>in</strong>formation from several<br />

levels of biotic organisation (<strong>in</strong>dividuals, populations,<br />

communities and systems) and focused on <strong>the</strong> dynamics<br />

of fl oodpla<strong>in</strong>–channel nutrients and <strong>the</strong> movement of<br />

larvae, juvenile and adult fi sh and <strong>the</strong>ir macro<strong>in</strong>vertebrate<br />

prey.<br />

The experiences from large schemes, especially <strong>the</strong><br />

KRRP, suggest <strong>the</strong>re are o<strong>the</strong>r reasons ra<strong>the</strong>r than a lack of<br />

ability as to why targets are not rout<strong>in</strong>ely set. In simple<br />

terms, it may be diffi cult, time consum<strong>in</strong>g and expensive.<br />

Particularly if <strong>the</strong> ‘ultimate’ approach to target sett<strong>in</strong>g,<br />

namely sett<strong>in</strong>g quantitative targets and formulat<strong>in</strong>g expectations<br />

as hypo<strong>the</strong>ses that can be evaluated statistically<br />

to provide objective evaluations, is pursued (Toth and<br />

Anderson, 1998). Such an approach requires an extensive<br />

<strong>in</strong>formation base on <strong>the</strong> river <strong>in</strong> question. Where this is<br />

absent, ga<strong>the</strong>r<strong>in</strong>g such data may not be cost effective, especially<br />

<strong>in</strong> <strong>the</strong> case of reach-scale projects, s<strong>in</strong>ce understand<strong>in</strong>g<br />

a complex system could easily take several years and<br />

exceed <strong>the</strong> lifetime of <strong>the</strong> project (Kondolf, 1998). As a<br />

result, a more pragmatic approach may be to set more<br />

qualitative targets, although it should be recognised that<br />

even qualitative targets may be evaluated <strong>in</strong> a rigorous<br />

manner. Toth and Anderson (1998) suggest that if qualitative<br />

expectations are to be used as success criteria <strong>the</strong>y must<br />

be expressed <strong>in</strong> an objective scientifi c manner relative to<br />

<strong>the</strong> restoration goal. Assessments of <strong>the</strong> presence or absence<br />

of species, for example, are highly sensitive to sampl<strong>in</strong>g<br />

effort (see Section 8.2.6) and an <strong>in</strong>creased frequency of<br />

occurrence may be a response that cannot be used to differentiate<br />

restoration from habitat enhancement.<br />

Fur<strong>the</strong>r pragmatism to target sett<strong>in</strong>g may be to simply<br />

narrow <strong>the</strong> targets (Toth and Anderson, 1998), such as<br />

focus<strong>in</strong>g on a s<strong>in</strong>gle species, as adopted <strong>in</strong> <strong>the</strong> recovery of<br />

endangered species populations. Many such species, from<br />

mammals (e.g. Eurasian otter Lutra lutra; Ott<strong>in</strong>o and<br />

Giller, 2004) to fi sh (e.g. Colorado pikem<strong>in</strong>now Ptychocheilus<br />

lucius <strong>in</strong> <strong>the</strong> <strong>River</strong> Colorado – van Steeter<br />

and Pitlick, 1998; Roanoke logperch Perc<strong>in</strong>a rex <strong>in</strong> <strong>the</strong><br />

Nottoway & Roanoke rivers <strong>in</strong> Virg<strong>in</strong>ia – Rosenberger and<br />

Angermeier, 2003) to <strong>in</strong>vertebrates (e.g. freshwater pearl<br />

mussel Margaritifera margaritifera – Cosgrove and Hastie,<br />

2001; White-clawed crayfi sh Austropotamobius pallipes<br />

– Smith et al., 1996) have been subject to <strong>in</strong>-depth assessment<br />

of habitat requirements <strong>in</strong> river<strong>in</strong>e environments.<br />

After detailed research, specifi c habitat attributes may be<br />

readily defi ned for particular life history stages.<br />

Defi n<strong>in</strong>g habitat relationships between any organism<br />

and any particular physical parameter, especially fl ow, has


<strong>Uncerta<strong>in</strong>ty</strong> Surround<strong>in</strong>g <strong>the</strong> Ecological Targets and Response of <strong>River</strong> and Stream <strong>Restoration</strong> 145<br />

been exploited by <strong>the</strong> Instream Flow Incremental Methodology<br />

(IFIM) (Gore and Judy, 1981; Bovee, 1982) and its<br />

underly<strong>in</strong>g model, <strong>the</strong> Physical Habitat Simulation Model<br />

(PHABSIM) (see Appendix 8.1). PHABSIM provides a<br />

potentially powerful tool that may enable specifi c targets<br />

to be set for particular species where enough <strong>in</strong>formation<br />

of <strong>the</strong> habitat relationships of <strong>the</strong> species concerned exists.<br />

Specifi c sets of habitat relationships, which may feed <strong>in</strong>to<br />

restoration efforts, are particularly well advanced for charismatic<br />

and/or commercially valuable species such as salmonid<br />

fi shes, particularly <strong>in</strong> <strong>the</strong> United States, where<br />

detailed and structured attempts at relatively large scale<br />

habitat restoration for salmonids has been go<strong>in</strong>g on for<br />

40 years or more (White & Brynildson, 1967; Wesche,<br />

1985) with a wealth of detailed guides and manuals on<br />

<strong>the</strong> web (e.g. http://www.wildfi sh.montana.edu/resources/<br />

mammals.asp).<br />

In some cases <strong>the</strong> use of <strong>in</strong>dicators such as fi sh (or birds<br />

or mammals) at <strong>the</strong> top of <strong>the</strong> food web may at least <strong>in</strong>dicate<br />

<strong>the</strong> direction and strength of <strong>the</strong> response of lower<br />

trophic levels as a result of habitat restoration. This has<br />

<strong>the</strong> basic assumption that targeted improvements <strong>in</strong> higher<br />

trophic levels as a result of improvements <strong>in</strong> habitat diversity<br />

are also likely to have led to a response <strong>in</strong> <strong>the</strong> diversity<br />

of lower trophic levels (i.e. species diversity is related to<br />

habitat diversity; Gorman and Karr, 1978). Monitor<strong>in</strong>g<br />

keystone species may not always be appropriate, however,<br />

as it may not provide enough <strong>in</strong>formation about o<strong>the</strong>r<br />

important processes such as organic matter spiral<strong>in</strong>g and<br />

ground–surface water <strong>in</strong>teractions (Tockner and Schiemer,<br />

1997).<br />

Never<strong>the</strong>less, where multiple measures are taken, <strong>the</strong><br />

multiple expectations generated may perversely lead to an<br />

ambiguous evaluation of restoration success if all expectations<br />

are not met (Toth and Anderson, 1998). In such a<br />

case, ei<strong>the</strong>r all expectations need to be judged collectively,<br />

for example by track<strong>in</strong>g a cumulative count of expectations<br />

that are achieved over time, or perhaps expectations<br />

could also be weighted accord<strong>in</strong>g to set priorities. Targets<br />

that can be expressed as constants or thresholds are<br />

perhaps easiest to evaluate but probably will be limited to<br />

a few attributes, for example <strong>in</strong> many systems enough<br />

<strong>in</strong>formation is available to formulate constant or threshold<br />

expectations for species richness. This approach is readily<br />

employed for <strong>in</strong>vertebrates, with <strong>the</strong> recent development<br />

of numerous assessment criteria particularly <strong>in</strong> relation to<br />

fl ow at <strong>the</strong> assemblage level (Appendix 8.2).<br />

Due to <strong>the</strong> natural spatial and temporal variability that<br />

is associated with <strong>the</strong> scale of ecosystem restoration, Toth<br />

and Anderson (1998) suggest many targets need to be<br />

expressed as ranges, although <strong>the</strong> breadth of <strong>the</strong> range that<br />

is used for restoration expectations cannot be so great as<br />

to mask or preclude evaluation of restoration success. For<br />

example, <strong>the</strong> expectation of small fi sh densities of 0.7<br />

to 1.3 fi sh per square foot of restored marshes on <strong>the</strong><br />

Kissimmee fl oodpla<strong>in</strong> refl ected sampl<strong>in</strong>g variability whilst<br />

still provid<strong>in</strong>g a useful range for evaluat<strong>in</strong>g restoration<br />

success. With h<strong>in</strong>dsight, our own experiences on small<br />

rivers <strong>in</strong> <strong>the</strong> United K<strong>in</strong>gdom show that such targets could<br />

have been readily set. For example, at several sites on <strong>the</strong><br />

Misbourne, <strong>the</strong> fi sh community responded <strong>in</strong> a ra<strong>the</strong>r predictable<br />

way to <strong>the</strong> <strong>in</strong>crease <strong>in</strong> fl ow, with one assemblage<br />

replaced by ano<strong>the</strong>r as habitat conditions changed (Appendix<br />

8.3).<br />

Target sett<strong>in</strong>g <strong>in</strong> large and complex schemes, especially<br />

<strong>the</strong> KRRP, but not <strong>in</strong> smaller schemes is <strong>the</strong> <strong>in</strong>verse of that<br />

expected. Perhaps it is simply that sett<strong>in</strong>g mean<strong>in</strong>gful<br />

targets and goals may be diffi cult, time consum<strong>in</strong>g and<br />

expensive. In large schemes <strong>in</strong> which huge <strong>in</strong>vestment of<br />

time and resources is made, it is prudent that every effort<br />

is made to reduce <strong>the</strong> uncerta<strong>in</strong>ty of <strong>the</strong> outcome, particularly<br />

where this relates to reducible ignorance <strong>in</strong> <strong>the</strong> form<br />

of <strong>in</strong>creas<strong>in</strong>g <strong>the</strong> knowledge base. Conversely, smaller<br />

schemes, where <strong>the</strong> risks of failure may not be as great and<br />

where resources are fewer and tend to be directed at ‘do<strong>in</strong>g<br />

restoration’, may embrace all aspects of uncerta<strong>in</strong>ty.<br />

In general, contrary to <strong>the</strong> conclusions of Downs and<br />

Sk<strong>in</strong>ner (2002), we suggest that ecologists may set targets<br />

as readily and as well as eng<strong>in</strong>eers given <strong>the</strong> opportunity,<br />

especially s<strong>in</strong>ce Stewardson and Ru<strong>the</strong>rfurd (Chapter 5)<br />

argue that <strong>the</strong>re may be unreasonable confi dence <strong>in</strong> <strong>the</strong><br />

target sett<strong>in</strong>g ability of geomorphologists, who tend to rely<br />

on uncalibrated (with real fi eld data) <strong>the</strong>oretical models<br />

dogged by poor estimation of variance.<br />

8.2.6 Time Scales and Trajectories<br />

In any restoration, consideration should be given to <strong>the</strong><br />

time scale for recovery or restoration, which is <strong>in</strong>fl uenced<br />

by how far removed <strong>the</strong> system is from <strong>the</strong> goal and targets<br />

set, and <strong>the</strong> trajectory and pathway(s) that may be undertaken<br />

(Gregory and Downs, Chapter 13). Anand and<br />

Desrochers (2004) observe that <strong>the</strong> lack of long term<br />

observational studies <strong>in</strong> virtually any ecological system<br />

often compels researchers to rely on <strong>the</strong>oretical models to<br />

speculate upon <strong>the</strong> trajectory that will lead <strong>the</strong> damaged<br />

system to <strong>the</strong> desired state. Accord<strong>in</strong>g to complex systems<br />

<strong>the</strong>ory, which <strong>in</strong>corporates concepts such as chaos, <strong>the</strong><br />

nature and direction of a trajectory is governed by <strong>the</strong><br />

system’s attractor, which may be thought of as its dest<strong>in</strong>ation.<br />

It is hoped that <strong>the</strong> dest<strong>in</strong>ation and desired state are<br />

one and <strong>the</strong> same th<strong>in</strong>g or <strong>the</strong> restoration attempt may be<br />

doomed at <strong>the</strong> outset. Anand and Desrochers (2004)<br />

eloquently illustrate <strong>the</strong> different types of attractor, with


146 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

<strong>the</strong> simplest be<strong>in</strong>g <strong>the</strong> progression of a system to <strong>the</strong> same<br />

stable state regardless of <strong>in</strong>itial conditions, analogous to a<br />

climax state of vegetational succession. Where a system<br />

has more than one attractor, it may cycle between two<br />

alternative states. However, systems may have any number<br />

of attract<strong>in</strong>g states and <strong>the</strong> system may progress to a particular<br />

‘bas<strong>in</strong> of attraction’ depend<strong>in</strong>g on <strong>the</strong> start<strong>in</strong>g po<strong>in</strong>t<br />

or <strong>the</strong> <strong>in</strong>itial conditions at which restoration started. This<br />

offers <strong>the</strong> possibility of multiple alternative stable states,<br />

which proved to be highly <strong>in</strong>fl uential <strong>in</strong> <strong>the</strong> restoration of<br />

shallow lakes (Figure 8.2, Scheffer et al., 1993). Ecological<br />

stability itself is also a diffi cult concept to grasp, as<br />

large spatial and temporal variation <strong>in</strong> populations of even<br />

key species tuned <strong>in</strong> to <strong>the</strong>ir life cycles or disturbance<br />

events (Figure 8.3) may mask underly<strong>in</strong>g trends. Stability<br />

is thus best judged over decades ra<strong>the</strong>r than from one year<br />

to <strong>the</strong> next.<br />

To <strong>the</strong> best of our knowledge such concepts have yet to<br />

be explored <strong>in</strong> rivers, although river ‘types’ are <strong>the</strong> basis<br />

of <strong>the</strong> Rosgen classifi cation used extensively and not<br />

without controversy <strong>in</strong> river restoration (Malakoff, 2004).<br />

Notable gaps (i.e. unquantifi ed uncerta<strong>in</strong>ties) <strong>in</strong> our conceptual<br />

understand<strong>in</strong>g of ecological dynamics with<strong>in</strong><br />

rivers appear to constra<strong>in</strong> any ability to predict how long<br />

nature will take to return <strong>the</strong> system to a specifi ed historic<br />

condition or even what is to be expected.<br />

Y<br />

X<br />

8.3 DEVISING A MONITORING PROGRAMME<br />

8.3.1 Basic Design<br />

To m<strong>in</strong>imise ecological uncerta<strong>in</strong>ty, monitor<strong>in</strong>g needs to<br />

be undertaken prior to <strong>the</strong> work <strong>in</strong> <strong>the</strong> form of collection<br />

of a robust set of basel<strong>in</strong>e data (see previously). If <strong>the</strong><br />

pre-restoration system is not fully understood, <strong>the</strong> results<br />

of <strong>the</strong> scheme cannot be correctly assessed or evaluated.<br />

Use of reference reaches (see previously) may help set<br />

targets and, to monitor progress aga<strong>in</strong>st those targets, a<br />

suitable monitor<strong>in</strong>g programme needs to be devised. Monitor<strong>in</strong>g<br />

of controls (e.g. unaltered reaches) undoubtedly<br />

helps reduce uncerta<strong>in</strong>ty, as this should help dist<strong>in</strong>guish<br />

what degree of change is due to natural <strong>in</strong>ter-annual variation<br />

and what is due to <strong>the</strong> restoration works (Stewardson<br />

and Ru<strong>the</strong>rfurd, Chapter 5).<br />

Choos<strong>in</strong>g which measurements of biotic and abiotic<br />

patterns and processes to monitor is <strong>the</strong>n critical, as <strong>the</strong>se<br />

need to be relevant to <strong>the</strong> goals of <strong>the</strong> project and <strong>the</strong><br />

targets set, and must be able to be l<strong>in</strong>ked to progress<br />

towards those targets. Where <strong>the</strong> target of restoration is a<br />

particular species, this may lead to very specifi c targets<br />

and <strong>the</strong> desire to monitor a few aspects. However, several<br />

workers have highlighted <strong>the</strong> need for wider surveillance<br />

ra<strong>the</strong>r than aim<strong>in</strong>g monitor<strong>in</strong>g at one aspect of <strong>the</strong> ecosystem.<br />

For example, Boon (1998) suggested that fi sheries-<br />

(Left) A po<strong>in</strong>t attractor. The same f<strong>in</strong>al state is reached <strong>in</strong> both <strong>in</strong>stances X and Y, although <strong>the</strong><br />

time taken to achieve this (shown by differ<strong>in</strong>g arrow lengths) varies depend<strong>in</strong>g on <strong>the</strong> location<br />

of <strong>the</strong> start<strong>in</strong>g position on <strong>the</strong> restoration trajectory.<br />

(Right) A dual attractor. This functions like a pendulum sw<strong>in</strong>g<strong>in</strong>g between two magnets, or<br />

alternative stable states (A and B). In this <strong>in</strong>stance <strong>the</strong> restoration time scales are <strong>the</strong> same,<br />

although <strong>the</strong> f<strong>in</strong>al state reached may depend on o<strong>the</strong>r ecological factors, for example composition<br />

of <strong>the</strong> community at <strong>the</strong> po<strong>in</strong>t when restoration is commenced.<br />

Figure 8.2 Types of attractors (Reproduced from M. Anand and R. Desrochers (2004) ‘Quantifi cation of <strong>Restoration</strong> Success Us<strong>in</strong>g<br />

Complex Systems Concepts and Models’ <strong>Restoration</strong> Ecology 12 (1), 117–123, with k<strong>in</strong>d permisison from Blackwell Publish<strong>in</strong>g.)<br />

Y<br />

X


<strong>Uncerta<strong>in</strong>ty</strong> Surround<strong>in</strong>g <strong>the</strong> Ecological Targets and Response of <strong>River</strong> and Stream <strong>Restoration</strong> 147<br />

(A) The potential effects of disturbance events (d) on ecosystem structure over time.<br />

Disturbance events could constitute anyth<strong>in</strong>g from pollution or alien species <strong>in</strong>troduction to <strong>the</strong><br />

impacts of restoration e.g. dredg<strong>in</strong>g, and recovery times differ accord<strong>in</strong>gly.<br />

(B) Variation <strong>in</strong> flow at two temporal scales.<br />

Figure 8.3 The effects of extensive spatial and temporal scale variation on ecological stability (Reproduced from P. S. White and<br />

J. L. Walker (1997) ‘Approximat<strong>in</strong>g nature’s variation: select<strong>in</strong>g and us<strong>in</strong>g reference <strong>in</strong>formation <strong>in</strong> restoration ecology’ <strong>Restoration</strong><br />

Ecology 5, 338–349, with k<strong>in</strong>d permission from Blackwell Publish<strong>in</strong>g.)<br />

led restoration projects should take a comprehensive<br />

ecosystem approach. Tockner et al. (1998) reached a<br />

similar conclusion ‘A key challenge <strong>in</strong> <strong>the</strong> evaluation of<br />

<strong>the</strong> effects of restoration is <strong>the</strong> development and test<strong>in</strong>g of<br />

an appropriate monitor<strong>in</strong>g scheme, which has to <strong>in</strong>clude<br />

a wide range of physical, chemical, geomorphic, and ecological<br />

parameters.’ Monitor<strong>in</strong>g a range of variables provides<br />

more <strong>in</strong>formation about <strong>the</strong> system response and <strong>the</strong><br />

potential mechanisms beh<strong>in</strong>d <strong>the</strong> response of <strong>the</strong> target<br />

variable or species.<br />

Consequently, a number of groups across <strong>the</strong> variety of<br />

trophic levels are typically used as <strong>in</strong>dicators of ecological<br />

status and response. For example, benthic <strong>in</strong>vertebrates,<br />

fi sh and aquatic macrophytes were selected as <strong>in</strong>dicators<br />

of <strong>the</strong> ecological status of rivers <strong>in</strong> <strong>the</strong> European<br />

Union Water Framework Directive (EU, 2000), and benthic


148 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

<strong>in</strong>vertebrates, fi sh, plankton, birds, amphibians and terrestrial<br />

and aquatic plants have been used on <strong>the</strong> project<br />

on <strong>the</strong> Austrian Danube (Buijse et al., 2002). For evaluat<strong>in</strong>g<br />

biodiversity of river–fl oodpla<strong>in</strong> complexes, a number<br />

of species-specifi c groups that differ <strong>in</strong> <strong>the</strong>ir responses to<br />

hydrological connectivity, water quality and habitat heterogeneity<br />

should be considered (Buijse et al., 2002;<br />

Tockner et al., 1999). The absence of certa<strong>in</strong> species may<br />

also be a good <strong>in</strong>dication of ecological deterioration (EU,<br />

2000). For example, <strong>the</strong> loss of long-distance migratory<br />

fi sh (salmonids, coregonids, shads and sturgeons) <strong>in</strong>dicates<br />

disruption of longitud<strong>in</strong>al connectivity or a deterioration<br />

of spawn<strong>in</strong>g/nursery areas.<br />

8.3.2 Sampl<strong>in</strong>g Requirements<br />

Brooks et al. (2002) po<strong>in</strong>ted out that many commonly<br />

used river restoration techniques may never have been<br />

scientifi cally tested and, <strong>in</strong> a demonstration of <strong>the</strong> need<br />

for rigorous scientifi c test<strong>in</strong>g, carried out a fi eld experiment<br />

designed to mimic restoration to <strong>in</strong>vestigate <strong>the</strong><br />

importance of habitat heterogeneity for macro<strong>in</strong>vertebrates.<br />

The results showed that although a diversity of<br />

habitats was required, macro<strong>in</strong>vertebrate populations were<br />

actually more sensitive to <strong>in</strong>dividual site conditions at<br />

each riffl e than to <strong>the</strong> heterogeneity treatments. This led<br />

to <strong>the</strong> conclusion that <strong>the</strong> extremely high variability<br />

between replicate riffl es meant that a monitor<strong>in</strong>g programme<br />

for localised restoration projects would be<br />

unlikely to detect gradual shifts <strong>in</strong> community structure<br />

until <strong>the</strong> differences between <strong>the</strong> reference and treatment<br />

sites were extreme. Brooks et al. <strong>the</strong>n suggested that <strong>in</strong>novative<br />

measurement of o<strong>the</strong>r parameters, such as ecosystem<br />

function variables (e.g. production, respiration,<br />

decomposition), may be more appropriate <strong>in</strong>dicators of<br />

change at local scales.<br />

This illustrates <strong>the</strong> age-old problem faced by ecologists<br />

of huge spatial and temporal variation <strong>in</strong> <strong>the</strong>ir subject<br />

matter (Figure 8.3). However, ra<strong>the</strong>r than look<strong>in</strong>g to o<strong>the</strong>r<br />

variables that may be technically diffi cult to measure,<br />

parameterise and ultimately <strong>in</strong>terpret, <strong>the</strong> only option to<br />

overcome variability is to devise appropriate sampl<strong>in</strong>g<br />

regimes of suffi cient <strong>in</strong>tensity and frequency and to use<br />

robust methods. In <strong>the</strong> case of <strong>in</strong>vertebrates, methods that<br />

allow changes to be evaluated on a community ra<strong>the</strong>r than<br />

species-specifi c level are likely to be useful (Appendix<br />

8.2). For fi sh, Bohl<strong>in</strong> et al. (1990) suggested three precision<br />

classes for fi sheries studies depend<strong>in</strong>g on <strong>the</strong> nature<br />

of <strong>the</strong> change that needs to be detected, i.e. a factor as<br />

small as 1.2, 1.5 or as large as 2.0, with about 80% probability<br />

when us<strong>in</strong>g a 5% signifi cance level. Unfortunately,<br />

when <strong>the</strong> number of fi sh is relatively low with even moder-<br />

ate variation around a mean value, even <strong>the</strong> lowest precision<br />

class tends to require a high sampl<strong>in</strong>g effort. On <strong>the</strong><br />

<strong>River</strong> Lambourn, UK, dur<strong>in</strong>g an attempt to monitor <strong>the</strong><br />

status of bullhead Cottus gobio, one of <strong>the</strong> species for<br />

which <strong>the</strong> river was designated as a site of <strong>in</strong>ternational<br />

conservation <strong>in</strong>terest (candidate Special Area of Conservation<br />

[cSAC]), it was calculated that twenty-seven 100 m<br />

sites would need to be surveyed over <strong>the</strong> 21 km river<br />

to detect change on a population level (Perrow and<br />

Toml<strong>in</strong>son, 2002). Us<strong>in</strong>g <strong>the</strong> same calculations to estimate<br />

<strong>the</strong> number of po<strong>in</strong>ts required <strong>in</strong> PASE (see Appendix 8.3)<br />

at each site, as few as 39 po<strong>in</strong>ts were required where <strong>the</strong><br />

population of bullheads was dense, but up to 150 po<strong>in</strong>ts<br />

were required where bullheads were at low density. The<br />

latter probably exceeds <strong>the</strong> number of po<strong>in</strong>ts that can be<br />

<strong>in</strong>dependently sampled, <strong>in</strong> that <strong>the</strong> po<strong>in</strong>ts are far enough<br />

apart to not <strong>in</strong>fl uence each o<strong>the</strong>r, at any particular site.<br />

Where such factors cannot be calculated <strong>in</strong> advance, rules<br />

of thumb are often generated and <strong>in</strong> <strong>the</strong> case of PASE this<br />

is often regarded as 50 po<strong>in</strong>ts (Garner, 1997). Below this,<br />

<strong>the</strong> confi dence <strong>in</strong> <strong>the</strong> estimates may be reduced, mak<strong>in</strong>g<br />

<strong>the</strong> response of <strong>the</strong> fi sh community to restoration diffi cult,<br />

if not impossible, to evaluate and <strong>in</strong>terpret. Too few<br />

samples may have been ano<strong>the</strong>r factor contribut<strong>in</strong>g to <strong>the</strong><br />

lack of a signifi cant impact of <strong>the</strong> <strong>in</strong>stallation of channel<br />

enhancements (i.e. artifi cial riffl es and fl ow defl ectors)<br />

upon fi sh abundance and species richness <strong>in</strong> low energy<br />

lowland alluvial channels (Pretty et al., 2003).<br />

The use of standardised methods and techniques <strong>in</strong> a<br />

‘one size fi ts all approach’ may also be a particular<br />

problem. Standardised approaches are particularly liked<br />

by statutory organisations and whilst <strong>the</strong> desire for comparability<br />

is understandable, many of <strong>the</strong> standard<br />

approaches have been designed for objectives o<strong>the</strong>r than<br />

monitor<strong>in</strong>g <strong>the</strong> response of a restoration scheme. Ultimately,<br />

such approaches may simply provide qualitative<br />

targets with no basis for statistical comparison. In relation<br />

to <strong>the</strong> recovery enhancement of <strong>the</strong> <strong>River</strong> Misbourne, UK<br />

(Appendix 8.3), standard <strong>River</strong> Corridor Survey (RCS),<br />

<strong>River</strong> Habitat Survey (RHS) and <strong>River</strong> Macrophyte Survey<br />

(RMS) were used at each site at not <strong>in</strong>considerable effort.<br />

However, all <strong>the</strong>se methods simply described and mapped<br />

<strong>the</strong> nature of <strong>the</strong> vegetation present and did not rout<strong>in</strong>ely<br />

provide quantitative measures that could be evaluated statistically.<br />

With <strong>the</strong> likely difference <strong>in</strong> hydrological regime<br />

between <strong>the</strong> different monitor<strong>in</strong>g sites along <strong>the</strong> course of<br />

<strong>the</strong> river, <strong>the</strong> response of each site had to be evaluated<br />

<strong>in</strong>dependently. Thus, <strong>the</strong> use of <strong>the</strong> standard fi sheries<br />

survey through depletion fi sh<strong>in</strong>g result<strong>in</strong>g <strong>in</strong> n = 1 on each<br />

sampl<strong>in</strong>g occasion could provide no more than a subjective<br />

comparison over time as fl ow was recovered. For this<br />

reason, a more specifi c sampl<strong>in</strong>g regime us<strong>in</strong>g PASE and


<strong>Uncerta<strong>in</strong>ty</strong> Surround<strong>in</strong>g <strong>the</strong> Ecological Targets and Response of <strong>River</strong> and Stream <strong>Restoration</strong> 149<br />

number of replicate samples with<strong>in</strong> each site was also used<br />

(Appendix 8.3).<br />

Even where <strong>the</strong> sampl<strong>in</strong>g regime has been carefully and<br />

perhaps specifi cally developed, particularly amongst fi sheries<br />

studies, it is essential to be clear that <strong>the</strong> effects of<br />

<strong>the</strong> scheme are at a population level with an <strong>in</strong>crease (or<br />

decrease) <strong>in</strong> <strong>the</strong> population, and not simply a change <strong>in</strong><br />

distribution of <strong>the</strong> fi sh present. Whilst <strong>the</strong> latter may still<br />

be viewed as a benefi cial response, as it illustrates that <strong>the</strong><br />

restoration works have produced <strong>the</strong> required sort of<br />

habitat, this may not have actually achieved <strong>the</strong> target<br />

set.<br />

8.3.3 Evaluation and Deviations from<br />

Expected Outcomes<br />

Put simply, <strong>the</strong> measure of success of any scheme is<br />

whe<strong>the</strong>r <strong>the</strong> targets have been met after a specifi ed time.<br />

Even where <strong>the</strong> plann<strong>in</strong>g and science were sound, poor<br />

implementation may have jeopardized <strong>the</strong> success of a<br />

restoration project. For example, <strong>in</strong> 1991 an attempt was<br />

made to restore vegetation on mud fl ats border<strong>in</strong>g <strong>the</strong><br />

Milwaukee <strong>River</strong>, USA, which had been previously submerged<br />

due to a dam (Drezner, 2004). Although <strong>the</strong> area<br />

was <strong>in</strong>itially seeded with native plants, and non-native<br />

plants were regularly pulled up, an oversight result<strong>in</strong>g <strong>in</strong><br />

<strong>the</strong> failure to remove Reed canary grass Phalaris arund<strong>in</strong>acea<br />

resulted <strong>in</strong> its dom<strong>in</strong>ation <strong>in</strong> certa<strong>in</strong> areas. When<br />

<strong>the</strong> area was surveyed 11 years later, on average 89% of<br />

plant stems per quadrat were non-native. The reasons<br />

beh<strong>in</strong>d such problems are not necessarily simple carelessness.<br />

Frequently, budgetary decisions relat<strong>in</strong>g to a whole<br />

project have to be made at <strong>the</strong> plann<strong>in</strong>g stage, which<br />

means that <strong>the</strong> fi nancial fl exibility required to deal with<br />

additional and unpredictable implementation issues may<br />

simply not be available.<br />

Streams and rivers have an <strong>in</strong>herent ability for natural<br />

recovery, which has been widely exploited <strong>in</strong> restoration<br />

(Brooks and Shields, 1996; Downs and Sk<strong>in</strong>ner, 2002).<br />

Consequently, <strong>the</strong>re seems to be little concern of whe<strong>the</strong>r<br />

a river will be restored once <strong>the</strong> limit<strong>in</strong>g factors are<br />

removed but ra<strong>the</strong>r when. Thus, once habitat restoration<br />

has taken place, ecological communities are left to recolonise.<br />

The rate of recovery follow<strong>in</strong>g a disturbance is<br />

dependent upon <strong>the</strong> resilience of <strong>the</strong> community to <strong>the</strong><br />

perturbation <strong>in</strong> <strong>the</strong> fi rst place as well as <strong>the</strong> rate of colonisation.<br />

Many <strong>in</strong>vertebrate species are highly resilient to<br />

disturbance through physical adaptations and behavioural<br />

change. Magoulik and Kobza (2003) concluded that many<br />

seek refuge from disturbance and/or have adaptations that<br />

provide refuge, whilst Townsend (1989) highlighted <strong>the</strong><br />

critical role played by refugia as sources of recolonisation<br />

after spates, and <strong>the</strong>refore as buffers aga<strong>in</strong>st disturbance.<br />

Important refugia <strong>in</strong>clude low fl ow areas (W<strong>in</strong>terbottom<br />

et al., 1997).<br />

The extent and <strong>in</strong>tensity of channel modifi cation (i.e.<br />

disturbance) is thought to have an impact on <strong>the</strong> time scale<br />

of recovery. Tikkanen et al. (1994) found that recolonisation<br />

of <strong>in</strong>vertebrates occurred rapidly, with<strong>in</strong> ten days,<br />

after a small-scale rehabilitation scheme. In contrast, on<br />

<strong>the</strong> <strong>River</strong> Rib where major habitat reconstruction was<br />

undertaken, although <strong>the</strong>re was rapid colonisation by<br />

some macro<strong>in</strong>vertebrates, <strong>the</strong> community took almost two<br />

years before show<strong>in</strong>g <strong>in</strong>dications of stabilisation. Fur<strong>the</strong>rmore,<br />

Niemi et al. (1990) found that many systems took<br />

more than fi ve years to recover to <strong>the</strong> desired ‘endpo<strong>in</strong>ts’,<br />

and <strong>the</strong>y concluded that <strong>the</strong> longest recovery times are<br />

associated with disturbance that leads to long term alterations<br />

<strong>in</strong> physical habitat. For example, <strong>the</strong> effects of an<br />

episodic pollution event are relatively short lived whilst<br />

recovery follow<strong>in</strong>g dredg<strong>in</strong>g takes much longer s<strong>in</strong>ce geomorphic<br />

channel readjustment is often slow. Indeed work<br />

carried out by Laasonon et al. (1998) on <strong>the</strong> recovery of<br />

macro<strong>in</strong>vertebrates follow<strong>in</strong>g stream habitat restoration<br />

works – which took <strong>the</strong> form of boulder dams, fl ow defl ectors,<br />

excavations and channel enlargements – <strong>in</strong>dicates<br />

that even after 16 years abundances of <strong>in</strong>vertebrates were<br />

still lower than <strong>in</strong> natural streams. Abundances of shredders<br />

were particularly low <strong>in</strong> recently restored streams <strong>in</strong><br />

comparison to streams restored a number of years previously,<br />

although even some streams that had been restored<br />

8 or 16 years ago still conta<strong>in</strong>ed relatively sparse shredder<br />

populations. Fur<strong>the</strong>r work (Muotka et al., 2002) highlighted<br />

<strong>the</strong> importance of aquatic mosses, which are frequently<br />

uprooted dur<strong>in</strong>g restoration works, and which<br />

perform important ecosystem functions, e.g. <strong>the</strong> retention<br />

of fi ne particles and <strong>the</strong> provision of fl ow refugia for<br />

<strong>in</strong>vertebrates, which ultimately <strong>in</strong>fl uence <strong>the</strong> rate of recovery<br />

of <strong>the</strong> macro<strong>in</strong>vertebrate community. At present <strong>the</strong><br />

recovery rates of aquatic mosses are unknown, mak<strong>in</strong>g it<br />

diffi cult to predict <strong>the</strong> responses of macro<strong>in</strong>vertebrate<br />

communities <strong>in</strong> <strong>the</strong>se types of systems.<br />

The time scales implicated <strong>in</strong> <strong>the</strong>se studies <strong>in</strong>dicate that<br />

all too often schemes are probably assessed too quickly,<br />

at a po<strong>in</strong>t dur<strong>in</strong>g <strong>the</strong> colonisation process ra<strong>the</strong>r that at <strong>the</strong><br />

end po<strong>in</strong>t target, even if this was defi ned accurately <strong>in</strong> <strong>the</strong><br />

fi rst place. Short term fl uctuations <strong>in</strong> species populations<br />

are to be expected before a longer-term stable equilibrium<br />

is reached.<br />

There is a wide range of factors <strong>in</strong>fl uenc<strong>in</strong>g <strong>the</strong> rate of<br />

recolonsiation that may simply over-ride <strong>the</strong> level of disturbance<br />

undertaken, <strong>in</strong>clud<strong>in</strong>g <strong>the</strong> source of colonists<br />

with<strong>in</strong> <strong>the</strong> river itself, <strong>the</strong> proximity of watercourses<br />

nearby for aerial colonsiation by w<strong>in</strong>ged adults and <strong>the</strong>


150 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

life cycle of <strong>the</strong> <strong>in</strong>vertebrates concerned and <strong>the</strong>ir habitat<br />

preferences. On <strong>the</strong> Rib, <strong>the</strong>re were marked differences <strong>in</strong><br />

<strong>the</strong> rates of recovery of different species and four groups<br />

were readily recognised: early colonists, medium term<br />

colonists, seasonal colonists and long term colonists<br />

(Figure 8.4). The Olive dun mayfl y Baetis rhodani was a<br />

typical early colonist be<strong>in</strong>g an active swimmer as well as<br />

a typical component of <strong>the</strong> drift. Whilst <strong>the</strong> Riffl e beetle<br />

Elmis aenea also used <strong>the</strong> drift, as a resident of fast fl ow<strong>in</strong>g<br />

areas it is adapted to be<strong>in</strong>g washed away and thus tends to<br />

colonise more slowly. Blackfl y larvae Simulium spp. are<br />

typically abundant <strong>in</strong> <strong>the</strong> spr<strong>in</strong>g and only tend to colonise<br />

newly available areas at this time. A typical long term<br />

colonist <strong>in</strong> this system was <strong>the</strong> predatory caseless caddis<br />

Rhyacophila dorsalis, which relies on <strong>the</strong> adult w<strong>in</strong>ged<br />

stage to colonise, as <strong>the</strong> larvae are adapted to fast fl ows<br />

Figure 8.4 The response of <strong>the</strong> mayfl y Baetis rhodani and <strong>the</strong> riffl e beetle Elmis aenea at two re-wetted sites on <strong>the</strong> <strong>River</strong> Rib, UK.


<strong>Uncerta<strong>in</strong>ty</strong> Surround<strong>in</strong>g <strong>the</strong> Ecological Targets and Response of <strong>River</strong> and Stream <strong>Restoration</strong> 151<br />

and are unable to tolerate slower fl ows between suitable<br />

habitat patches.<br />

Like <strong>in</strong>vertebrates, fi sh are often assumed to be capable<br />

of rapid colonisation and, aga<strong>in</strong> like <strong>in</strong>vertebrates, <strong>the</strong><br />

speed and recovery of colonisation may be primarily<br />

dependent on <strong>the</strong> geographical isolation of <strong>the</strong> communities<br />

and <strong>the</strong> proximity of potential colonisers. Both downstream<br />

and upstream movement may be readily undertaken<br />

provided <strong>the</strong>re are no barriers to <strong>the</strong> latter. What constitutes<br />

a barrier varies enormously. Bullhead, a small<br />

benthic fi sh is effectively limited by barriers of just 10 cm<br />

(Utz<strong>in</strong>ger et al., 1998), whereas Barbel Barbus barbus a<br />

large (to 1 m) powerful cypr<strong>in</strong>id <strong>in</strong> European rivers has<br />

been recorded mak<strong>in</strong>g movements of up to nearly 20 km,<br />

cross<strong>in</strong>g large weirs (to 2 m) <strong>in</strong> <strong>the</strong> process (Lucas and<br />

Bately, 1996). The ability of many anadromous salmonids<br />

(salmon and trout) to leap large obstacles is well known.<br />

Even though <strong>the</strong> restored area may have been recolonised,<br />

recruitment success can still fl uctuate ow<strong>in</strong>g to a wide<br />

range of factors not related to <strong>the</strong> scheme (e.g. temperature,<br />

predation of eggs and larvae etc). Such factors can<br />

affect both <strong>the</strong> abundance and distribution of <strong>in</strong>dividuals,<br />

and may signifi cantly slow <strong>the</strong> colonisation process.<br />

Moreover, <strong>the</strong> major structur<strong>in</strong>g forces of predation and<br />

competition may rage for some time before a stable confi<br />

guration is reached. In <strong>the</strong> Misbourne for example, sticklebacks<br />

may have out-competed o<strong>the</strong>r species for some<br />

time, even as physical conditions changed, simply as <strong>the</strong>y<br />

arrived fi rst and achieved high density as a result of <strong>the</strong><br />

abundance of potential nest sites and fl ow refugia <strong>in</strong> <strong>the</strong><br />

form of emergent and <strong>the</strong>n submerged vegetation. Only<br />

when <strong>the</strong>se were removed, did <strong>the</strong> community shift rapidly,<br />

with bullheads dom<strong>in</strong>at<strong>in</strong>g through rapid recruitment<br />

(Appendix 8.3).<br />

In general, vagaries <strong>in</strong> <strong>the</strong> abilities of different organisms<br />

to colonise <strong>in</strong>tuitively mean that this stage of <strong>the</strong><br />

project is highly vulnerable to <strong>the</strong> <strong>in</strong>fl uence of ecological<br />

uncerta<strong>in</strong>ty. Predict<strong>in</strong>g <strong>the</strong> likely pattern of colonisation<br />

may be problematic, due to both a limited understand<strong>in</strong>g<br />

of o<strong>the</strong>r populations <strong>in</strong> <strong>the</strong> local area, and a general lack<br />

of scientifi c knowledge. For example, <strong>the</strong> factors <strong>in</strong>fl uenc<strong>in</strong>g<br />

<strong>the</strong> colonisation of submerged macrophytes, which are<br />

of enormous ecological importance, are poorly understood.<br />

Clearly, better knowledge of <strong>the</strong> colonisation mechanisms<br />

and movement patterns is vital to understand <strong>the</strong><br />

processes beh<strong>in</strong>d restoration of river<strong>in</strong>e systems (Fenoglio<br />

et al., 2002) and allow a better understand<strong>in</strong>g of <strong>the</strong> time<br />

scale required for recovery to a stable end po<strong>in</strong>t. A lack of<br />

scientifi c knowledge is not necessarily easy to overcome,<br />

s<strong>in</strong>ce <strong>the</strong> situation at one site may differ substantially from<br />

ano<strong>the</strong>r and at present <strong>the</strong>re is a lack of longer term studies<br />

that may feed <strong>in</strong>to <strong>the</strong> knowledge base (Figure 8.1).<br />

8.3.4 Appraisal and Feedback<br />

The importance of appraisal has been highlighted by<br />

Anderson and Dugger (1998) who concluded that ‘Failure<br />

to evaluate projects not only precludes learn<strong>in</strong>g anyth<strong>in</strong>g<br />

about a particular restoration, but it also limits <strong>the</strong> opportunity<br />

for improv<strong>in</strong>g plans for future projects’. In our experience<br />

<strong>in</strong> <strong>the</strong> United K<strong>in</strong>gdom <strong>the</strong> number of projects that<br />

<strong>in</strong>corporate post-project appraisal (PPA) is <strong>in</strong>creas<strong>in</strong>g follow<strong>in</strong>g<br />

early efforts to raise <strong>the</strong> importance of this aspect<br />

(ECON, 1993). Review<strong>in</strong>g projects from 1991–1996,<br />

Holmes (1998) recorded that 53% of projects had undergone<br />

PPA, 23% of projects did not <strong>in</strong>corporate any form<br />

of PPA and for 25% of projects this <strong>in</strong>formation was not<br />

known. Today, partly as a result of <strong>the</strong> efforts of <strong>the</strong> RRP<br />

demonstration projects and subsequently <strong>the</strong> <strong>River</strong> <strong>Restoration</strong><br />

Centre (RRC) with<strong>in</strong> <strong>the</strong> European Centre for <strong>River</strong><br />

<strong>Restoration</strong> (see previously), we would suggest that perhaps<br />

only <strong>the</strong> smaller scale schemes do not rout<strong>in</strong>ely undertake<br />

PPA. As well as <strong>the</strong> importance of shared <strong>in</strong>formation,<br />

drivers have <strong>in</strong>cluded <strong>the</strong> need for organisations such as<br />

<strong>the</strong> Environment Agency to demonstrate cost effi ciency<br />

for <strong>the</strong>ir ‘bus<strong>in</strong>ess’ needs, and that <strong>the</strong> publicity mach<strong>in</strong>es<br />

of <strong>the</strong> organisations <strong>in</strong>volved like/need to have someth<strong>in</strong>g<br />

to show to <strong>the</strong> public and o<strong>the</strong>r <strong>in</strong>terested parties.<br />

However, <strong>the</strong>re is potentially little control over <strong>the</strong><br />

quality of <strong>the</strong> <strong>in</strong>formation produced, with accumulat<strong>in</strong>g<br />

ecological uncerta<strong>in</strong>ties <strong>in</strong> <strong>the</strong> earlier stages of <strong>the</strong> project<br />

potentially ultimately result<strong>in</strong>g <strong>in</strong> mislead<strong>in</strong>g <strong>in</strong>formation<br />

that could be used to <strong>in</strong>form future projects. This is a<br />

variant of <strong>the</strong> ‘Almost as bad as no evaluation are poorly<br />

planned efforts that waste limited resources while provid<strong>in</strong>g<br />

mean<strong>in</strong>gless or even mislead<strong>in</strong>g <strong>in</strong>formation’ warned<br />

by Anderson and Dugger (1998). It is diffi cult if not<br />

impossible to assess how much this type of situation<br />

occurs. Although ecological uncerta<strong>in</strong>ty may frequently<br />

occur, <strong>the</strong> concern is that it goes unrecognised, which<br />

could lead to <strong>in</strong>accurate results and mis<strong>in</strong>formation. If<br />

ecological uncerta<strong>in</strong>ty is recognised, and if <strong>the</strong> scheme<br />

results <strong>in</strong> a clear improvement, even if it is not exactly as<br />

predicted, this is perhaps less of a concern. However, <strong>the</strong>re<br />

is <strong>the</strong> possibility that <strong>the</strong> project may be held <strong>in</strong> less<br />

esteem or considered not as successful if targets are not<br />

met, and project fund<strong>in</strong>g and <strong>the</strong> prospect of cont<strong>in</strong>ued<br />

restoration may be jeopardised, even though <strong>the</strong> work may<br />

have ultimately had a positive impact on <strong>the</strong> ecology of<br />

<strong>the</strong> site. Schemes are beg<strong>in</strong>n<strong>in</strong>g to allow for this eventuality<br />

and are constructed to allow adaptive management if<br />

<strong>the</strong> scheme needs to be adjusted to obta<strong>in</strong> more desirable<br />

results.<br />

Even if appraisal is conducted, it is often not easy to<br />

evaluate success and for this reason <strong>the</strong>re is a major move


152 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

from river ecologists to establish a set of criteria for measur<strong>in</strong>g<br />

<strong>the</strong> ecological success of schemes (Palmer et al.,<br />

2005). It is hoped that this will fi nally be endorsed by <strong>the</strong><br />

United Nations Environmental Programme. The criteria<br />

specifi ed are that:<br />

<strong>the</strong> design should be based on a specifi ed guid<strong>in</strong>g image<br />

that a more dynamic, healthy river could exist at <strong>the</strong><br />

site;<br />

<strong>the</strong> river’s ecological condition must be measurably<br />

improved;<br />

<strong>the</strong> river system must be more self-susta<strong>in</strong><strong>in</strong>g and resilient<br />

to external perturbations so that only m<strong>in</strong>imal<br />

follow-up ma<strong>in</strong>tenance is needed.<br />

dur<strong>in</strong>g <strong>the</strong> construction phase, no last<strong>in</strong>g harm should<br />

be <strong>in</strong>fl icted on <strong>the</strong> ecosystem.<br />

both pre- and post-assessment must be completed and<br />

data made publicly available.<br />

The emphasis beh<strong>in</strong>d <strong>the</strong>se criteria is <strong>the</strong> desire to have<br />

projects that are funded and implemented <strong>in</strong> <strong>the</strong> name of<br />

ecological restoration evaluated accord<strong>in</strong>g to ecological<br />

<strong>in</strong>dicators of success ra<strong>the</strong>r than accord<strong>in</strong>g to o<strong>the</strong>r social,<br />

economic and <strong>in</strong>stitutional factors, such as cost effectiveness,<br />

stakeholder satisfaction and aes<strong>the</strong>tic/recreational<br />

value (Palmer et al., 2005). These criteria are fl exible<br />

enough to apply to both large scale major projects and<br />

smaller projects where complex and expensive design is<br />

not appropriate. They also encourage <strong>the</strong> view of restoration<br />

success as an adaptive process ra<strong>the</strong>r than a s<strong>in</strong>gle<br />

defi ned end po<strong>in</strong>t.<br />

8.4 CONCLUSIONS AND RECOMMENDATIONS<br />

Throughout this chapter we have sought to outl<strong>in</strong>e sources<br />

of unwanted uncerta<strong>in</strong>ty that ultimately reduce confi dence<br />

<strong>in</strong> restoration and h<strong>in</strong>der <strong>the</strong> development of a cost effective,<br />

repeatable approach. The latter can only be achieved<br />

by contribut<strong>in</strong>g to a shared knowledge base, <strong>the</strong> potential<br />

for which is now enormous with a number of dedicated<br />

restoration journals and even more specifi c websites. The<br />

approach adopted accepts that much uncerta<strong>in</strong>ty result<strong>in</strong>g<br />

from natural variation and sheer system complexity, especially<br />

<strong>in</strong> large systems, will rema<strong>in</strong> unknowable. Fortunately,<br />

this has not dampened <strong>the</strong> desire for restoration of<br />

large systems, which make greatest contribution to biodiversity,<br />

as recent <strong>in</strong>itiatives with<strong>in</strong> <strong>the</strong> European Union to<br />

restore rivers such as <strong>the</strong> Danube (Tockner and Schiemer,<br />

1997), <strong>the</strong> Rh<strong>in</strong>e (Buijse et al., 2002) and <strong>the</strong> Drava and<br />

Lech rivers <strong>in</strong> Austria (Mohl, 2004) testify.<br />

Indeed, <strong>in</strong>vestment <strong>in</strong> large systems such as <strong>the</strong> KRRP<br />

has illustrated that many uncerta<strong>in</strong>ties may be quantifi ed<br />

and overcome, with <strong>the</strong> accumulation of an extensive<br />

knowledge and basel<strong>in</strong>e data, ga<strong>the</strong>red over a number of<br />

years, which <strong>in</strong> turn required considerable resources. With<br />

such effort, predictions and targets can be detailed, quantitative<br />

and set to with<strong>in</strong> a realistic time scale. This quells<br />

<strong>the</strong> notion of <strong>the</strong> <strong>in</strong>ability of ecologists to set targets which<br />

has been raised as a particular source of uncerta<strong>in</strong>ty. Our<br />

experiences, especially <strong>in</strong> <strong>the</strong> United K<strong>in</strong>gdom, suggest<br />

that <strong>the</strong> lack of target sett<strong>in</strong>g cannot be attributed to a<br />

particular <strong>in</strong>ability to do so but ra<strong>the</strong>r appears to be more<br />

due to a lack of resources <strong>in</strong> small projects especially.<br />

With an <strong>in</strong>crease <strong>in</strong> post-project appraisal of whe<strong>the</strong>r <strong>the</strong><br />

project succeeded or failed, it is more diffi cult to justify<br />

not us<strong>in</strong>g targets at all. Project ‘culture’ thus appears to<br />

be chang<strong>in</strong>g for <strong>the</strong> better.<br />

There is probably a better knowledge base of <strong>the</strong> fauna–<br />

habitat <strong>in</strong>formation than is immediately apparent and this<br />

may be developed specifi cally where this is lack<strong>in</strong>g,<br />

although this may be expensive of time and resources.<br />

Habitat relationships have, after all, been a fundamental<br />

component of many species-centred restoration projects,<br />

<strong>in</strong>clud<strong>in</strong>g <strong>in</strong> rivers. Powerful tools such as PHABSIM and<br />

a wide range of classifi cation systems are available to<br />

predict and quantify change even <strong>in</strong> species-rich groups<br />

such as <strong>in</strong>vertebrates, although of course <strong>the</strong> limitations<br />

of such tools do need to be recognised.<br />

There seems to be historical sense of ecologists be<strong>in</strong>g<br />

advocates of a ‘black art’, perhaps particularly to hardened<br />

river eng<strong>in</strong>eers deal<strong>in</strong>g <strong>in</strong> ma<strong>the</strong>matical formulae. This<br />

may have orig<strong>in</strong>ated from ecologists be<strong>in</strong>g employed to<br />

count ‘bugs’ or fi sh or record plant abundance, orig<strong>in</strong>at<strong>in</strong>g<br />

from eng<strong>in</strong>eer<strong>in</strong>g works. Ecology is essentially a numerate<br />

science that can be quantifi ed by relationships and patterns<br />

and is no more ‘uncerta<strong>in</strong>’ than any o<strong>the</strong>r science, although<br />

it does deal with extremely complex systems and <strong>the</strong> relationships<br />

and <strong>in</strong>teractions with<strong>in</strong> <strong>the</strong>m. Ecological understand<strong>in</strong>g<br />

also often takes time and resources. Br<strong>in</strong>g<strong>in</strong>g<br />

ecologists <strong>in</strong>to a more central role, with <strong>in</strong>volvement from<br />

<strong>the</strong> early stages of <strong>the</strong> project, may help reduce uncerta<strong>in</strong>ty.<br />

Engag<strong>in</strong>g a greater range of academic river ecologists<br />

with wide understand<strong>in</strong>g of catchment processes,<br />

<strong>the</strong>oreticians and ecological modellers, as well as <strong>the</strong><br />

‘samplers and sorters’, who are frequently consultants, is<br />

also thought likely to be particularly benefi cial.<br />

Even a lack of historical reference data may be partly<br />

alleviated by monitor<strong>in</strong>g reference (control) reaches before<br />

and after <strong>the</strong> impact of <strong>the</strong> restoration scheme <strong>in</strong> a before–<br />

after–control–impact (BACI) design (e.g. as on <strong>the</strong> <strong>River</strong>s<br />

Cole and Brede; biggs et al., 1998). Particularly where this<br />

<strong>in</strong>corporates any historical <strong>in</strong>formation, it lends itself well<br />

to target sett<strong>in</strong>g with <strong>the</strong> subsequent means of assess<strong>in</strong>g<br />

whe<strong>the</strong>r <strong>the</strong> target <strong>in</strong> <strong>the</strong> controls is reached. As more


<strong>Uncerta<strong>in</strong>ty</strong> Surround<strong>in</strong>g <strong>the</strong> Ecological Targets and Response of <strong>River</strong> and Stream <strong>Restoration</strong> 153<br />

<strong>in</strong>formation is ga<strong>the</strong>red, a fl exible adaptive management<br />

approach allows targets to be adjusted (or perhaps even<br />

hardened from qualitative to quantitative, if necessary),<br />

enables fur<strong>the</strong>r studies to be planned when needed and can<br />

recognise and deal with emerg<strong>in</strong>g problems.<br />

The rigorous selection or even design of appropriate<br />

monitor<strong>in</strong>g techniques is essential if <strong>the</strong> major source<br />

of uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> monitor<strong>in</strong>g phase is to be reduced.<br />

The use of standard techniques simply because <strong>the</strong>y are<br />

rout<strong>in</strong>ely used by <strong>the</strong> organisation <strong>in</strong>volved, is to be<br />

guarded aga<strong>in</strong>st, as such methods have often been developed<br />

for a different purpose. It is far better to treat each<br />

scheme as a unique <strong>in</strong>dividual experiment for which <strong>the</strong><br />

appropriate method, <strong>in</strong>tensity and frequency of sampl<strong>in</strong>g<br />

is calculated. In o<strong>the</strong>r words, monitor<strong>in</strong>g should be seen<br />

as detailed research ra<strong>the</strong>r than a mere estimation of<br />

general changes.<br />

Although rivers have enormous resilience and an<br />

<strong>in</strong>herent ability for natural recovery <strong>the</strong> time scale and<br />

trajectory of recovery/restoration represents a major<br />

source of uncerta<strong>in</strong>ty, particularly where a stable endpo<strong>in</strong>t<br />

has been poorly defi ned or <strong>the</strong>re is little understand<strong>in</strong>g of<br />

what that may be. The ‘how long is a piece of str<strong>in</strong>g’<br />

analogy does not sit well with project budgets, which are<br />

typically fi nite and allocated to a particular time frame.<br />

We suggest that <strong>the</strong> trajectory of restoration may be more<br />

direct and shortened by considered project selection <strong>in</strong> <strong>the</strong><br />

fi rst place, where restoration is part of a generic <strong>the</strong>me<br />

adopted by a statutory body or <strong>in</strong>terested organisation.<br />

Tackl<strong>in</strong>g less damaged sites with good prospects for<br />

recovery, particularly where this can be undertaken through<br />

recovery enhancement and not major habitat reconstruction,<br />

is also likely to lead to a reduced time scale for<br />

recovery.<br />

There is clearly a need for more work on <strong>the</strong> factors<br />

surround<strong>in</strong>g colonisation to enable better predictions to be<br />

ultimately made. This may need to be conducted specifi -<br />

cally for <strong>the</strong> river concerned with<strong>in</strong> <strong>the</strong> project <strong>in</strong> an adaptive<br />

manner and may also require action to speed up <strong>the</strong><br />

colonisation process. For example, with problem species<br />

such as <strong>in</strong>vertebrates with particularly poor dispersal abilities,<br />

it may be prudent to <strong>in</strong>troduce <strong>the</strong>se, perhaps through<br />

‘seed<strong>in</strong>g’ of substrate, to reach a stable confi guration as<br />

soon as possible. For fi sh, it is important to defi ne any<br />

barriers to natural movement that aga<strong>in</strong>, may be readily<br />

overcome by translocation.<br />

Obviously, <strong>the</strong> strength of post-project appraisal ultimately<br />

depends on <strong>the</strong> uncerta<strong>in</strong>ties accumulated earlier<br />

<strong>in</strong> <strong>the</strong> project. Where <strong>the</strong>se are great this could deliver a<br />

fl awed evaluation of little or no value to future projects.<br />

Follow<strong>in</strong>g <strong>the</strong> approach outl<strong>in</strong>ed it should be possible to<br />

avoid this scenario <strong>in</strong> <strong>the</strong> future.<br />

ACKNOWLEDGEMENTS<br />

We are grateful to <strong>the</strong> Environment Agency for fund<strong>in</strong>g<br />

several of <strong>the</strong> studies documented and to its staff for assistance<br />

with sampl<strong>in</strong>g. ECON staff undertook much of <strong>the</strong><br />

work on <strong>the</strong> <strong>River</strong> Misbourne.<br />

REFERENCES<br />

Adams W. 1997. Rationalization and conservation: Ecology and<br />

<strong>the</strong> management of nature <strong>in</strong> <strong>the</strong> UK. Transactions of <strong>the</strong><br />

Institute of British Geographers 22: 227–291.<br />

Anand M, Desrochers R. 2004. Quantifi cation of restoration<br />

success us<strong>in</strong>g complex systems concepts and models. <strong>Restoration</strong><br />

Ecology 12: 117–123.<br />

Anderson DH, Dugger BD. 1998. A conceptual basis for evaluat<strong>in</strong>g<br />

restoration success. Transactions of <strong>the</strong> North American<br />

Wildlife and Natural Resources Conference 63.<br />

Benndorf J. 1992. The control of <strong>in</strong>direct effects of biomanipulation.<br />

In: Sutcliffe DW, Jones JG (Eds), Eutrophication:<br />

Research and Application to Water Supply. Freshwater Biological<br />

Association: Cumbria, UK; 82–93.<br />

Biggs J, Cornfi eld A, Grøn R et al. 1998. <strong>Restoration</strong> of <strong>the</strong> rivers<br />

Brede, Cole and Skerne: a jo<strong>in</strong>t Danish and British EU–LIFE<br />

demonstration project, V – Short–term impacts on <strong>the</strong> conservation<br />

value of aquatic macro<strong>in</strong>vertebrate and macrophyte<br />

assemblages. Aquatic Conservation: Mar<strong>in</strong>e and Freshwater<br />

Ecosystems 8: 241–255.<br />

Bohl<strong>in</strong> T. 1990. Estimation of population parameters us<strong>in</strong>g electric<br />

fi sh<strong>in</strong>g: aspects of <strong>the</strong> sampl<strong>in</strong>g design with emphasis on<br />

salmonids <strong>in</strong> streams. In: Cowx IG, Lamarque P (Eds), Fish<strong>in</strong>g<br />

with Electricity Fish<strong>in</strong>g News Books: Oxford, UK; 156–173.<br />

Boon PJ. 1998. <strong>River</strong> restoration <strong>in</strong> fi ve dimensions. Aquatic<br />

Conservation: Mar<strong>in</strong>e and Freshwater Ecosystems 8:<br />

257–264.<br />

Bovee KD. 1982. A Guide to Stream Habitat Analysis Us<strong>in</strong>g <strong>the</strong><br />

Instream Flow Incremental Methodology. Instream Flow Information<br />

Paper No.12. US Fish and Wildlife Service, FES/OBS–<br />

82/86: Wash<strong>in</strong>gton, DC.<br />

Bradshaw AD. 1987. <strong>Restoration</strong>: an acid test for ecology. In:<br />

Jordan WR, Gilp<strong>in</strong> ME, Aber JD (Eds), <strong>Restoration</strong> Ecology<br />

Cambridge University Press: Cambridge, UK; 23–29.<br />

Brooks A, Shields FD Jr. 1996. Towards an approach to susta<strong>in</strong>able<br />

river restoration. In: Brookes A, Shields FD Jr., <strong>River</strong><br />

Channel <strong>Restoration</strong>: Guid<strong>in</strong>g Pr<strong>in</strong>ciples for Susta<strong>in</strong>able Projects.<br />

John Wiley & Sons Ltd: Chichester, UK; 385–402.<br />

Brooks SS, Palmer MA, Card<strong>in</strong>ale BJ et al. 2002. Assess<strong>in</strong>g<br />

stream ecosystem rehabilitation: Limitations of community<br />

structure. <strong>Restoration</strong> Ecology 10: 156–168.<br />

Bruce-Burgess L, Sk<strong>in</strong>ner L. 2002. Appraisal: <strong>River</strong> restoration’s<br />

miss<strong>in</strong>g l<strong>in</strong>k. Proceed<strong>in</strong>gs of Conference held 27th November<br />

2002 at <strong>the</strong> University of Nott<strong>in</strong>gham, Downloadable from<br />

http://www.<strong>the</strong>rrc.co.uk.<br />

Buijse AD, Coups H, Staras M et al. 2002. <strong>Restoration</strong> strategies<br />

for river fl oodpla<strong>in</strong>s along large lowland rivers <strong>in</strong> Europe.<br />

Freshwater Biology 47: 889–907.


154 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Calow P. 1992. Energy budgets. In: Calow P, Petts G (Eds), The<br />

<strong>River</strong>s Handbook, Volume 1. Blackwell Scientifi c Publications:<br />

Oxford, UK; 370–378.<br />

Carpenter SR, Kitchell JF. 1993. The Trophic Cascade <strong>in</strong> Lakes.<br />

Cambridge University Press: Cambridge, UK.<br />

Carle FL, Strub MR. 1978. A new method for estimat<strong>in</strong>g population<br />

size from removal data. Biometrics 34: 621–630.<br />

Chandler JR. 1970. A biological approach to water quality management.<br />

Water Pollution Control, 69: 415–422.<br />

Clarke SJ, Bruce-Burgess L, Wharton G. 2003. L<strong>in</strong>k<strong>in</strong>g form and<br />

function: towards an ecohycromorphic approach to susta<strong>in</strong>able<br />

river restoration. Aquatic Conservation: Mar<strong>in</strong>e and Freshwater<br />

Ecosystems. 13: 439–450.<br />

Copp GH, Peñáz 1988. Ecology of fi sh spawn<strong>in</strong>g and nursery<br />

zones <strong>in</strong> <strong>the</strong> fl ood pla<strong>in</strong>, us<strong>in</strong>g a new sampl<strong>in</strong>g approach.<br />

Hydrobiologia 169: 209–224.<br />

Cosgrove PJ, Hastie LC. 2001. Conservation of threatened freshwater<br />

pearl mussel population: river management, mussel<br />

translocation and confl ict resolution. Biological Conservation<br />

99: 183–190.<br />

de Waal LC, Large ARG, Gippel CJ, Wade PM. 1995. <strong>River</strong> and<br />

fl oodpla<strong>in</strong> restoration <strong>in</strong> western Europe: opportunities and<br />

constra<strong>in</strong>ts. Archiv für Hydrobiologie 101: 679–693.<br />

Downs P, Sk<strong>in</strong>ner KS. 2002. <strong>River</strong>s and streams. In: Perrow MR,<br />

Davy AJ (Eds), Handbook of Ecological <strong>Restoration</strong>. Volume<br />

2: <strong>Restoration</strong> <strong>in</strong> Practice. Cambridge University Press: Cambridge,<br />

UK; 267–296.<br />

Drezner TD. 2004. Few native species colonize on mud fl ats ten<br />

years after dam removal. <strong>Restoration</strong> Ecology 22: 50–51.<br />

ECON 1993. The <strong>River</strong> <strong>Restoration</strong> Project, Phase 1: <strong>the</strong> Feasibility<br />

Study. Report to <strong>River</strong> <strong>Restoration</strong> Project (RRP), UK.<br />

Extence CA, Balbi DM, Chadd RP. 1999. <strong>River</strong> fl ow <strong>in</strong>dex<strong>in</strong>g<br />

us<strong>in</strong>g British benthic macro-<strong>in</strong>vertebrates: a framework for<br />

sett<strong>in</strong>g hydroecological objectives. Regulated <strong>River</strong>s Research<br />

and Management 15: 543–574.<br />

EU. 2000. Directive 2000/60/EC of <strong>the</strong> European Parliament<br />

and of <strong>the</strong> council establish<strong>in</strong>g framework for <strong>the</strong> community<br />

action <strong>in</strong> <strong>the</strong> fi eld of water policy. (Website: http://europa.eu.<br />

<strong>in</strong>t/comm/environment/water/water-framework/<strong>in</strong>dex_en.<br />

html).<br />

Fenolglio S, Agosta P, Bo T, Cucco M. 2002. Field experiments<br />

on colonization and movements of stream <strong>in</strong>vertebrates <strong>in</strong> an<br />

Apenn<strong>in</strong>e river (Visone, NW Italy). Hydrobiologia 474:<br />

125–130.<br />

Gameson ALH, Wheeler A. 1977. <strong>Restoration</strong> and <strong>the</strong> recovery of<br />

<strong>the</strong> Thames Estuary. In: Cairns J Jr., Dickson KL, Herricks EE<br />

(Eds), Recovery and <strong>Restoration</strong> <strong>in</strong> Damaged Ecosystems<br />

University Press of Virg<strong>in</strong>ia: Charlottesville, Virg<strong>in</strong>ia; 72–101.<br />

Garner P. 1997. Sample sizes for length and density estimation<br />

of 0+ fi sh when us<strong>in</strong>g po<strong>in</strong>t sampl<strong>in</strong>g by electrofi sh<strong>in</strong>g. Journal<br />

of Fisheries Biology 50: 95–106.<br />

Gore SA, Judy RD Jr. 1981. Predictive models of benthic macro<strong>in</strong>vertebrate<br />

density for use <strong>in</strong> <strong>in</strong>stream fl ow studies and<br />

regulated fl ow management. Canadian Journal of Fisheries<br />

and Aquatic Sciences 38: 1363–1370.<br />

Gorman OT, Karr JR. 1978. Habitat structure and stream fi sh<br />

communities. Ecology 59 (3): 507–515.<br />

Harper DM, Smith CD, Barham PJ. 1992. Habitats as <strong>the</strong> build<strong>in</strong>g<br />

blocks for river conservation assessment. In: Boon PJ, Calow<br />

P, Petts GE (Eds), <strong>River</strong> Conservation and Management. John<br />

Wiley & Sons Ltd: Chichester, UK; 311–319.<br />

Hearne J, Johnson IW, Armitage PD. 1994. Determ<strong>in</strong>ation of<br />

ecologically acceptable fl ows <strong>in</strong> rivers with seasonal changes<br />

<strong>in</strong> <strong>the</strong> density of macrophytes. Regulated <strong>River</strong>s: Research and<br />

Management 9: 117–184.<br />

Hoffmann CC, Pedersen ML, Kronvang B, Øvig L. 1998. <strong>Restoration</strong><br />

of <strong>the</strong> <strong>River</strong>s Brede, Cole and Skerne: a jo<strong>in</strong>t Danish and<br />

British EU–LIFE demonstration project, IV – implications for<br />

nitrate and iron transformation. Aquatic Conservation: Mar<strong>in</strong>e<br />

and Freshwater Ecosystems 8: 223–240.<br />

Holmes NTH. 1998. A Review of <strong>River</strong> Rehabilitation <strong>in</strong> <strong>the</strong> UK,<br />

1990–1996. Technical Report to <strong>the</strong> Environment Agency<br />

(W175), Environment Agency: Bristol, UK.<br />

Holmes NTH, Nielsen MB. 1998. <strong>Restoration</strong> of <strong>the</strong> rivers Brede,<br />

Cole and Skerne: a jo<strong>in</strong>t Danish and British EU–LIFE demonstration<br />

project, I – Sett<strong>in</strong>g up and delivery of <strong>the</strong> project.<br />

Aquatic Conservation: Mar<strong>in</strong>e and Freshwater Ecosystems 8:<br />

185–196.<br />

Jeppesen E, Sammalkarpi I. 2002. Lakes In: Perrow MR, Davy<br />

AJ (Eds), Handbook of Ecological <strong>Restoration</strong>. Volume 2: <strong>Restoration</strong><br />

<strong>in</strong> Practice. Cambridge University Press: Cambridge,<br />

UK; 297–324.<br />

Junk WJ, Bayley PB, Sparks RE. 1989. The fl ood pulse concept<br />

<strong>in</strong> river-fl oodpla<strong>in</strong> systems. Special Publication of <strong>the</strong> Canadian<br />

Journal of Fisheries and Aquatic Sciences 106:<br />

110–127.<br />

Kern K. 1992. Rehabilitation of streams <strong>in</strong> South-west Germany.<br />

In: Boon PJ, Calow P, Petts GE (Eds), <strong>River</strong> Conservation and<br />

Management. John Wiley & Sons Ltd: Chichester, UK;<br />

321–335.<br />

Kondolf GM. 1998. Lessons learned from river restoration projects<br />

<strong>in</strong> California. Aquatic Conservation: Mar<strong>in</strong>e and Freshwater<br />

Ecosystems 8: 39–52.<br />

Kronvang B, Svendsen LM, Brookes A et al. 1998. <strong>Restoration</strong><br />

of <strong>the</strong> rivers Brede, Cole and Skerne: a jo<strong>in</strong>t Danish and British<br />

EU–LIFE demonstration project, III – channel morphology,<br />

hydrodynamics and transport of sediment and nutrients.<br />

Aquatic Conservation: Mar<strong>in</strong>e and Freshwater Ecosystems 8:<br />

209–222.<br />

Laasonen P, Muotka T, Kilvijärvi I. 1998. Recovery of macro<strong>in</strong>vertebrate<br />

communities from stream habitat restoration.<br />

Aquatic Conservation: Mar<strong>in</strong>e and Freshwater Ecosystems 8:<br />

101–113.<br />

Lepori F, Palm D, Malmqvist B. 2005. Effects of stream restoration<br />

on ecosystem function<strong>in</strong>g: detritus retentiveness and<br />

decomposition. Journal of Applied Ecology 42: 175–189.<br />

Lucas MC, Batley E. 1996. Seasonal movements and behaviour<br />

of adult barbel Barbus barbus, a river<strong>in</strong>e cypr<strong>in</strong>id fi sh: implications<br />

for river management. Journal of Applied Ecology 33:<br />

1345–1358<br />

Magoulick DD, Kobza R. 2003. The role of refugia for fi shes<br />

dur<strong>in</strong>g drought: a review and syn<strong>the</strong>sis. Freshwater Biology 48:<br />

1186–1198.<br />

Malakoff D. 2004. The <strong>River</strong> Doctor. Science 305: 937–939.


<strong>Uncerta<strong>in</strong>ty</strong> Surround<strong>in</strong>g <strong>the</strong> Ecological Targets and Response of <strong>River</strong> and Stream <strong>Restoration</strong> 155<br />

McDonald A, Lane SN, Haycock NE, Chalk EA. 2004. <strong>River</strong>s of<br />

dreams: on <strong>the</strong> gulf between <strong>the</strong>oretical and practical aspects<br />

of an upland river restoration. Transactions of <strong>the</strong> Institute of<br />

British Geographers 29: 257–281.<br />

Milner NJ, Hemsworth RJ, Jones BE. 1985. Habitat evaluation<br />

as a fi sheries management tool. Journal of Fish Biology 27<br />

(Supplement A): 85–108.<br />

Mohl A. 2004. LIFE <strong>River</strong> <strong>Restoration</strong> projects <strong>in</strong> Austria. In:<br />

Geres D (Ed) <strong>River</strong> <strong>Restoration</strong> 2004 – Pr<strong>in</strong>ciples, Processes,<br />

Practices. Proceed<strong>in</strong>gs of <strong>the</strong> 3rd ECRR International Conference<br />

on <strong>River</strong> <strong>Restoration</strong> <strong>in</strong> Europe, Zagreb: Croatia (Papers<br />

downloadable from http://www.ecrr.org/).<br />

Muotka T, Paavol R, Haapala A et al. 2002. Long-term recovery<br />

of stream-habitat structure and benthic <strong>in</strong>vertebrate communities<br />

from <strong>in</strong>-stream restoration. Biological Conservation 105:<br />

243–253.<br />

Nestler JM, Schneider LT, Latka D, Johnson P. 1996. Impact analysis<br />

and restoration plann<strong>in</strong>g us<strong>in</strong>g <strong>the</strong> river<strong>in</strong>e community<br />

habitat assessment and restoration concept (RCHARC). In:<br />

Leclerc M, Capra H, Valent<strong>in</strong> S et al. (Eds), Ecohydraulics 2000.<br />

Proceed<strong>in</strong>gs of <strong>the</strong> 2nd International Symposium on Habitat<br />

Hydraulics, INRS – Eau: Québec, Canada; A871-A876.<br />

Newbold JD. 1992. Cycles and spirals of nutrients. In: Calow P,<br />

Petts GE (Eds), The <strong>River</strong>s Handbook Volume 1. Blackwell<br />

Scientifi c Publications: Oxford, UK; 79–408.<br />

Newson MD, Pitlick J, Sear DA. 2002. Runn<strong>in</strong>g water: fl uvial<br />

geomorphology and river restoration. In: Perrow MR, Davy AJ<br />

(Eds), Handbook of Ecological <strong>Restoration</strong>. Volume 1: Pr<strong>in</strong>ciples<br />

of <strong>Restoration</strong>. Cambridge University Press: Cambridge,<br />

UK: 324–354.<br />

Niemi GJ, DeVore P, Detenheck N et al. 1990. Overview of case<br />

studies on recovery of aquatic systems from disturbance.<br />

Environmental Management 14: 571–587.<br />

NRA (National <strong>River</strong>s Authority)/Kirkpatrick Scott Wilson.<br />

1992. Method for Assessment of Low Flow Conditions Caused<br />

by Abstraction. National <strong>River</strong>s Authority, R & D Note 45.<br />

Environment Agency: Bristol, UK.<br />

NRA (National <strong>River</strong>s Authority). 1995. Surface Water Abstraction<br />

Licens<strong>in</strong>g Policy Development. Project Report 505,<br />

National <strong>River</strong>s Authority: Bristol, UK.<br />

NRC (National Research Council US). 1992. Chapter 5: rivers<br />

and streams. In: <strong>Restoration</strong> of Aquatic Ecosytems National<br />

Academy Press: Wash<strong>in</strong>gton, DC.<br />

Ormerod SJ. 2004. A golden age of river restoration science?<br />

Aquatic Conservation: Mar<strong>in</strong>e & Freshwater Ecosystems. 14<br />

(6): 543–549.<br />

Ott<strong>in</strong>o P, Giller P. 2004. Distribution, density, diet and habitat use<br />

of <strong>the</strong> otter <strong>in</strong> relation to land use <strong>in</strong> <strong>the</strong> Aragl<strong>in</strong> Valley, Sou<strong>the</strong>rn<br />

Ireland. Biology and Environment: Proceed<strong>in</strong>gs of <strong>the</strong><br />

Royal Irish Academy 104B (1): 1–17.<br />

Palmer MA, Bernhardt ES, Allan JD et al. 2005. Standards for<br />

ecologically successful river restoration. Journal of Applied<br />

Ecology, 42 (2): 208–217.<br />

Perrow MR, Jowitt AJD, Zambrano González L. 1996. Sampl<strong>in</strong>g<br />

fi sh communities <strong>in</strong> shallow lowland lakes: po<strong>in</strong>t-sample electric<br />

fi sh<strong>in</strong>g vs electric fi sh<strong>in</strong>g with<strong>in</strong> stop-nets. Fisheries<br />

Management & Ecology 3: 303–313.<br />

Perrow MR, Toml<strong>in</strong>son ML, Zambrano L. 2002. Fish. In: Perrow<br />

MR, Davy AJ (Eds), Handbook of Ecological <strong>Restoration</strong>.<br />

Volume 1: Pr<strong>in</strong>ciples of <strong>Restoration</strong>. Cambridge University<br />

Press: Cambridge, UK; 324–354.<br />

Perrow MR, Davy AJ (Eds). 2002a. Handbook of Ecological<br />

<strong>Restoration</strong>. Volume 1: Pr<strong>in</strong>ciples of <strong>Restoration</strong>. Cambridge<br />

University Press: Cambridge, UK.<br />

Perrow MR, Davy AJ (Eds). 2002b. Handbook of Ecological<br />

<strong>Restoration</strong>. Volume 2: <strong>Restoration</strong> <strong>in</strong> Practice. Cambridge<br />

University Press: Cambridge, UK.<br />

Perrow MR, Toml<strong>in</strong>son ML. 2002. <strong>River</strong> Lambourn cSAC basel<strong>in</strong>e<br />

survey: Bullhead and Lamprey populations. Report to<br />

English Nature.<br />

Pretty JL, Harrison SSC, Shepherd DJ et al. 2003. <strong>River</strong> rehabilitation<br />

and fi sh populations: assess<strong>in</strong>g <strong>the</strong> benefi t of <strong>in</strong>stream<br />

structures. Journal of Applied Ecology 40: 251–265.<br />

Punchard NT, Perrow MR, Jowitt AJD. 2000. Fish habitat associations,<br />

community structure, density and biomass <strong>in</strong> natural and<br />

channelized lowland streams <strong>in</strong> <strong>the</strong> catchment of <strong>the</strong> <strong>River</strong><br />

Wensum, UK. In: Cowx IG (Ed.), Management and Ecology<br />

of <strong>River</strong> Fisheries. Blackwell Science: Oxford, UK; 143–157.<br />

Richardson JS, Jackson MJ. 2002. Aquatic <strong>in</strong>vertebrates. In:<br />

Perrow MR, Davy AJ (Eds), Handbook of Ecological <strong>Restoration</strong>.<br />

Volume 1: Pr<strong>in</strong>ciples of <strong>Restoration</strong>. Cambridge University<br />

Press: Cambridge, UK; 300–323.<br />

Rob<strong>in</strong>son C, Whitton S. 2004. Fisheries Action Plans: a new<br />

approach to public consultation and <strong>the</strong> impetus for habitat<br />

enhancement works <strong>in</strong> <strong>the</strong> upper Thames catchment (England).<br />

In: Geres D (Ed.), <strong>River</strong> <strong>Restoration</strong> 2004 – Pr<strong>in</strong>ciples, Processes,<br />

Practices. Proceed<strong>in</strong>gs of <strong>the</strong> 3rd ECRR International<br />

Conference on <strong>River</strong> <strong>Restoration</strong> <strong>in</strong> Europe: Zagreb, Croatia.<br />

Papers downloadable from http://www.ecrr.org.<br />

Roni P, Beechie TJ, Bilby RE et al. 2002. A review of stream<br />

restoration techniques and a hierarchical strategy for prioritiz<strong>in</strong>g<br />

restoration <strong>in</strong> <strong>the</strong> Pacifi c Northwest watersheds. North<br />

American Journal of Fisheries Management 22: 1–20.<br />

Rosenberger A, Angermeier PL. 2003. Ontogenetic shifts <strong>in</strong><br />

habitat use by <strong>the</strong> endangered Roanoke logperch (Perc<strong>in</strong>a rex).<br />

Freshwater Biology 48: 1563–1577.<br />

RSPB (Royal Society for <strong>the</strong> Protection of Birds), NRA (National<br />

<strong>River</strong>s Authority) and <strong>the</strong> RSNC (Royal Society for Nature<br />

Conservation). 1995. The New <strong>River</strong>s and Wildlife Handbook<br />

The Royal Society for <strong>the</strong> Protection of Birds: Sandy, Bedfordshire,<br />

UK.<br />

Russell IC. 1994. Salmon stock<strong>in</strong>g <strong>in</strong> north-east England – some<br />

factors affect<strong>in</strong>g return rates. In: Cowx IG (Ed.), Rehabilitation<br />

of Freshwater Fisheries. Fish<strong>in</strong>g News Books, Blackwell Scientifi<br />

c Publications Ltd: Oxford, UK; 255–267.<br />

Scheffer M, Hosper SH, Meijer M-L et al. 1993. Alternative<br />

equilibria <strong>in</strong> shallow lakes. Trends <strong>in</strong> Ecology and Evolution 8:<br />

275–279.<br />

Smith GRT, Learner MA, Slater FM, Foster J. 1996. Habitat features<br />

important for <strong>the</strong> conservation of <strong>the</strong> native crayfi sh <strong>in</strong><br />

Brita<strong>in</strong>. Biological Conservation 75: 239–246.<br />

Tikkanen P, Laasonen O, Muotka T et al. 1994. Short term recovery<br />

of benthos follow<strong>in</strong>g disturbance from stream habitat rehabilitation.<br />

Hydrobiologia 273: 121–130.


156 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Tockner K, Schiemer F. 1997. Ecological aspects of <strong>the</strong> restoration<br />

strategy for a river-fl oodpla<strong>in</strong> system on <strong>the</strong> Danube <strong>River</strong><br />

<strong>in</strong> Austria. Global Ecology and Biogeography Letters 6:<br />

321–329.<br />

Tockner K, Schiemer F, Ward JV. 1998. Conservation by restoration:<br />

<strong>the</strong> management concept for a river-fl oodpla<strong>in</strong> system on<br />

<strong>the</strong> Danube <strong>River</strong> <strong>in</strong> Austria. Aquatic Conservation 8: 71–86.<br />

Tockner K, Schiemer F, Baumgartner C et al. 1999. The Danube<br />

<strong>Restoration</strong> Project: Species diversity patterns across connectivity<br />

gradient6s <strong>in</strong> <strong>the</strong> fl oodpla<strong>in</strong> system. Regulated <strong>River</strong>s:<br />

Research and Management. 15: 245–258.<br />

Toth LA. 1996. Restor<strong>in</strong>g <strong>the</strong> hydrogeomorphology of <strong>the</strong> channelised<br />

Kissimmee <strong>River</strong>. In: Brookes A, Shields FD Jr (Eds),<br />

<strong>River</strong> Channel <strong>Restoration</strong>: Guid<strong>in</strong>g Pr<strong>in</strong>ciples for Susta<strong>in</strong>able<br />

Projects John Wiley & Sons Ltd: Chichester, UK; 369–383.<br />

Toth LA, Anderson DH. 1998. Develop<strong>in</strong>g expectations for ecosystem<br />

restoration. In: Transactions of <strong>the</strong> 63rd North American<br />

Wildlife Natural Resources Conference: Wash<strong>in</strong>gton, DC;<br />

122–134.<br />

Toth LA, Melv<strong>in</strong> SL, Arr<strong>in</strong>gton DA, Chamberla<strong>in</strong> J. 1998. Hydrological<br />

manipulations of <strong>the</strong> channelised Kissimmee <strong>River</strong>:<br />

Implications for restoration. Bioscience 48: 758–764.<br />

Townsend CR. 1989. The patch dynamics concept of stream community<br />

ecology. Journal of <strong>the</strong> North American Benthological<br />

Society 8: 36–50.<br />

Trexler JC. 1995. <strong>Restoration</strong> of <strong>the</strong> Kissimmee <strong>River</strong>: a conceptual<br />

model of past and present fi sh communities and its consequences<br />

for evaluat<strong>in</strong>g restoration success. <strong>Restoration</strong> Ecology<br />

3: 195–210.<br />

Utz<strong>in</strong>ger J, Roth C, Peter A. 1998. Effects of environmental<br />

parameters on <strong>the</strong> distribution of bullhead Cottus gobio with<br />

particular consideration of <strong>the</strong> effects of obstructions. Journal<br />

of Applied Ecology 35: 882–892.<br />

Van Steeter MM, Pitlick J. 1998. Geomorphology and endangered<br />

fi sh habitats of <strong>the</strong> Upper Colorado <strong>River</strong>. 1: Historic changes<br />

<strong>in</strong> streamfl ow, sediment load and channel morphology. Water<br />

Resources Research 34: 287–302.<br />

Vannote RL, M<strong>in</strong>shall GW, Cumm<strong>in</strong>s KW et al. 1980. The river<br />

cont<strong>in</strong>uum concept. Canadian Journal of Fisheries and Aquatic<br />

Sciences 37: 130–137.<br />

Vivash R, Ottosen O, Janes M, Sørensen HV. 1998. <strong>Restoration</strong><br />

of <strong>the</strong> rivers Brede, Cole and Skerne: a jo<strong>in</strong>t Danish and British<br />

EU–LIFE demonstration project, II – <strong>the</strong> river restoration<br />

works and o<strong>the</strong>r related practical aspects. Aquatic Conservation:<br />

Mar<strong>in</strong>e and Freshwater Ecosystems 8: 197–208.<br />

Waldman JR, Wirg<strong>in</strong> II. 1998. Status and restoration options for<br />

Atlantic sturgeon <strong>in</strong> North America. Conservation Biology 12:<br />

631–638.<br />

Wesche TA. 1985. Stream channel modifi cations and reclamation<br />

structures to enhance fi sh habitat. In: Gore JA (Ed.), The<br />

<strong>Restoration</strong> of <strong>River</strong>s and Streams: Theories and Experience.<br />

Butterworth Publishers: Massachusetts, USA; 103–163.<br />

Whalen PJ, Toth LA, Koebel JW, Strayer PK. 2002. Kissimmee<br />

<strong>River</strong> <strong>Restoration</strong>: A Case Study. Water Science and Technology<br />

45: 55–62.<br />

White RJ, Brynildson OM. 1967. Guidel<strong>in</strong>es for Management of<br />

Trout Stream Habitat <strong>in</strong> Wiscons<strong>in</strong>. Technical Bullet<strong>in</strong> No. 39,<br />

Division of Conservation, Department of Natural Resources:<br />

Madison, Wiscons<strong>in</strong>.<br />

White PS, Walker JL. 1997. Approximat<strong>in</strong>g nature’s variation:<br />

select<strong>in</strong>g and us<strong>in</strong>g reference <strong>in</strong>formation <strong>in</strong> restoration<br />

ecology. <strong>Restoration</strong> Ecology 5: 338–349.<br />

W<strong>in</strong>terbottom JH, Orton SE, Hildrew AG. 1997. Field Experiments<br />

on fl ow refugia <strong>in</strong> streams. Freshwater Biology 37:<br />

569–588.<br />

APPENDIX 8.1 THE USE AND POTENTIAL<br />

LIMITATIONS OF IFIM AND PHABSIM<br />

The ‘Instream fl ow <strong>in</strong>cremental methodology’ (IFIM) and<br />

its underly<strong>in</strong>g model, <strong>the</strong> ‘Physical habitat simulation<br />

model’ (PHABSIM) represent <strong>the</strong> most popular approach<br />

for simulat<strong>in</strong>g habitat preferences (Downs and Sk<strong>in</strong>ner,<br />

2002). The relationship between streamfl ow and available<br />

physical habitat (defi ned by depth, velocity, substrate and<br />

cover) is used to compute <strong>the</strong> ‘weighted usable area’<br />

(WUA) <strong>in</strong> a reach as a function of <strong>the</strong> river discharge, for<br />

different life stages and species of fi sh. For each life<br />

stage of <strong>the</strong> target species, <strong>the</strong> model requires expressions<br />

of <strong>the</strong> relative suitability for that species of <strong>the</strong> full<br />

range of values taken by <strong>the</strong>se variables. These univariate<br />

curves or habitat suitability <strong>in</strong>dices may be derived from<br />

exist<strong>in</strong>g literature, expert op<strong>in</strong>ion or by sampl<strong>in</strong>g techniques<br />

such as electro-fi sh<strong>in</strong>g or snorkell<strong>in</strong>g. PHABSIM<br />

also conta<strong>in</strong>s a number of hydraulic models that predict<br />

values of depth and velocity at different simulation discharges.<br />

These models require calibration us<strong>in</strong>g fi eld data<br />

collected at two or more calibration discharges. Observations<br />

of substrate and cover are recorded us<strong>in</strong>g a cod<strong>in</strong>g<br />

system and are assumed to be <strong>in</strong>dependent of discharge.<br />

Once calibrated <strong>the</strong> model can simulate values of microhabitat<br />

variables over <strong>the</strong> full range of discharge with<strong>in</strong> a<br />

river reach. Comb<strong>in</strong><strong>in</strong>g <strong>the</strong> results with habitat suitability<br />

data produces <strong>the</strong> WUA versus Discharge relationship<br />

(CEH website: http://www.nwl.ac.uk/ih). PHABSIM,<br />

by relat<strong>in</strong>g habitat to discharge, provides a quantitative<br />

entity, allow<strong>in</strong>g river ecologists to negotiate prescribed<br />

fl ows <strong>in</strong> equivalent terms to o<strong>the</strong>r water resource demands,<br />

and offers a practical means of <strong>in</strong>tegrat<strong>in</strong>g ecological<br />

requirements of aquatic species with o<strong>the</strong>r water resource<br />

demands.<br />

However, numerous procedural, biological and physical<br />

limit<strong>in</strong>g assumptions prevented IFIM and PHABSIM<br />

from be<strong>in</strong>g reliable and, <strong>in</strong> response to <strong>the</strong>se criticisms, a<br />

wide range of more complex approaches based on<br />

advanced understand<strong>in</strong>g of <strong>the</strong> bioenergetics of fi sh<br />

biology and on more realistic representation and modell<strong>in</strong>g<br />

of fl ow patterns is now at <strong>the</strong> research stage (3rd<br />

International Symposium on Ecohydraulics, 1999).


<strong>Uncerta<strong>in</strong>ty</strong> Surround<strong>in</strong>g <strong>the</strong> Ecological Targets and Response of <strong>River</strong> and Stream <strong>Restoration</strong> 157<br />

At <strong>the</strong> Instream Flow Incremental Methodology Workshop<br />

held <strong>in</strong> New Zealand (February 2004), jo<strong>in</strong>tly hosted<br />

by Fish and Game New Zealand and <strong>the</strong> Department of<br />

Conservation, Cawthron Institute NIWA (Conference<br />

website: http://www.doc.govt.nz/Explore/Hunt<strong>in</strong>g-and-<br />

Fish<strong>in</strong>g/Taupo-Fishery/), several speakers (e.g. Maclean<br />

and Death) expressed concern about some of <strong>the</strong> underly<strong>in</strong>g<br />

assumptions beh<strong>in</strong>d PHABSIM, <strong>in</strong>clud<strong>in</strong>g that: fi sh are<br />

limited by habitat; that habitat preferences are adequately<br />

described by depth, velocity and substrate type; that habitat<br />

preferences do not change <strong>in</strong> accordance with diurnal/seasonal<br />

patterns; and that a large area of suboptimal habitat<br />

is preferable to a small area of optimal habitat.<br />

It was also argued that PHABSIM was not appropriate<br />

for use <strong>in</strong> connection with New Zealand (NZ) <strong>in</strong>vertebrates<br />

on <strong>the</strong> basis that depth, velocity and substrate type<br />

were of m<strong>in</strong>or importance <strong>in</strong> comparison with o<strong>the</strong>r factors<br />

such as riparian catchment vegetation, disturbance and<br />

food supply. Habitat requirements also appeared to be<br />

fl exible for many NZ species. The use of habitat suitability<br />

curves was also brought <strong>in</strong>to question and it was noted<br />

that it needs to be made explicitly clear how <strong>the</strong>y were<br />

developed, for example which sites were used, <strong>the</strong> number<br />

of samples taken and whe<strong>the</strong>r diurnal patterns were<br />

accounted for <strong>in</strong> <strong>the</strong> sampl<strong>in</strong>g process. The lack of variance<br />

estimation (i.e. error bars) was also raised as a limitation.<br />

Reduced (or <strong>in</strong>creased) fl ows might have effects on<br />

o<strong>the</strong>r variables that are not taken <strong>in</strong>to account, such as<br />

nutrient content, sediment build up, and temperature.<br />

PHABSIM and variants such as <strong>the</strong> <strong>River</strong><strong>in</strong>e Community<br />

Habitat Assessment and <strong>Restoration</strong> Concept<br />

(RCHARC) (Nestler et al., 1996) <strong>in</strong> <strong>the</strong> United K<strong>in</strong>gdom<br />

on lowland chalk streams suffer<strong>in</strong>g from low fl ow have<br />

also proved to be of limited use for <strong>in</strong>vertebrates. There is<br />

<strong>the</strong> risk that habitat area/habitat response predictions will<br />

be mean<strong>in</strong>gless if <strong>the</strong> sampl<strong>in</strong>g transects are not widely<br />

representative. These problems may be more marked for<br />

smaller rivers and streams <strong>in</strong> <strong>the</strong> UK context. It is known<br />

that for chalk streams and o<strong>the</strong>r nutrient-rich lowland<br />

rivers <strong>the</strong> seasonal changes <strong>in</strong> <strong>the</strong> growth and biomass of<br />

macrophytes require, at <strong>the</strong> very least, a careful calibration<br />

of PHABSIM to prevent distortion of results (Hearne et<br />

al., 1994).<br />

Overall, whilst PHABSIM is a potentially powerful tool<br />

it may be <strong>the</strong> most effective when evaluat<strong>in</strong>g and compar<strong>in</strong>g<br />

different management options, <strong>in</strong> <strong>in</strong>stances where<br />

physical habitat limits populations. The United K<strong>in</strong>gdom<br />

method only assesses physical habitat, so factors such as<br />

water quality and temperature and sediment transport<br />

require complementary studies. Indeed, where such factors<br />

are <strong>the</strong> prime constra<strong>in</strong>t on populations <strong>the</strong> use of<br />

PHABSIM is <strong>in</strong>appropriate.<br />

Much criticism centres on this latter po<strong>in</strong>t and<br />

PHABSIM like any model is only as good as <strong>the</strong> data that<br />

is entered and makes <strong>the</strong> assumption that habitat variables<br />

are <strong>the</strong> best descriptor of <strong>the</strong> abundance of a particular<br />

group, which of course need not be <strong>the</strong> case. Also, <strong>the</strong><br />

quantity and quality of water cannot limit <strong>the</strong> distribution<br />

or abundance of <strong>the</strong> specifi ed organism <strong>in</strong> <strong>the</strong> fi rst place.<br />

Whilst <strong>the</strong>re is an <strong>in</strong>herent desire for scientists to adopt a<br />

quantitative approach <strong>in</strong> order to reduce uncerta<strong>in</strong>ty, this<br />

is only successful if a high level of confi dence can be<br />

placed <strong>in</strong> <strong>the</strong> predictions. In <strong>the</strong> case of PHABSIM, many<br />

users have found that predictions for even ‘straightforward’<br />

schemes do not come close to <strong>the</strong> quantitative<br />

targets set. However, PHABSIM can also be used as a<br />

qualitative tool, and may be useful when it comes to<br />

select<strong>in</strong>g <strong>the</strong> best likely restoration option from a number<br />

of possible schemes.<br />

APPENDIX 8.2 CLASSIFICATION TECHNIQUES<br />

FOR MONITORING AND POTENTIALLY<br />

PREDICTING THE RESPONSE OF<br />

INVERTEBRATES TO FLOW RESTORATION<br />

In <strong>the</strong> United Kongdom low fl ows are seen to be a problem,<br />

especially <strong>in</strong> chalk streams, which are characterised by<br />

high <strong>in</strong>vertebrate biomass and abundance and a relatively<br />

high diversity of species. The ma<strong>in</strong> physical effects of low<br />

fl ows <strong>in</strong> chalk streams are: long term dry<strong>in</strong>g of ephemeral<br />

or w<strong>in</strong>terbourne reaches; long term dry<strong>in</strong>g of spr<strong>in</strong>g seepages<br />

or fl ushes; downstream migration of <strong>the</strong> perennial<br />

head of a chalk stream or river; and reduced water levels<br />

and/or velocity <strong>in</strong> downstream reaches. Much restoration<br />

<strong>in</strong> <strong>the</strong> United K<strong>in</strong>gdom has been concerned with ‘Alleviation<br />

of Low Flow’ (ALF) schemes, which were driven<br />

by <strong>the</strong> concern of <strong>the</strong> environmental impacts of overabstraction.<br />

The advantages and disadvantages of methodologies<br />

developed to monitor <strong>in</strong>vertebrates <strong>in</strong> low fl ow<br />

systems are outl<strong>in</strong>ed below.<br />

Scott Wilson Kirkpatrick (SWK) Methodology<br />

The fi rst national methodology for assess<strong>in</strong>g low fl ow<br />

conditions caused by abstraction was not written until<br />

1992 (NRA R&D Note 45, 1992 – a procedural manual<br />

by <strong>the</strong> consultants Scott Wilson Kirkpatrick). This set out<br />

a po<strong>in</strong>t-scor<strong>in</strong>g prioritis<strong>in</strong>g procedure based on <strong>the</strong> assessment<br />

of four <strong>in</strong>dicators: Hydrological; Ecological; Landscape<br />

and Amenity; and Public Perception.<br />

For each <strong>in</strong>dicator, <strong>in</strong>dividual assessments of a variety<br />

of parameters were conducted. The scores for each <strong>in</strong>dicator<br />

were comb<strong>in</strong>ed, with separate weight<strong>in</strong>gs. The ecological<br />

<strong>in</strong>dicator parameters and <strong>the</strong>ir <strong>in</strong>dividual weight<strong>in</strong>gs


158 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

were: <strong>in</strong>vertebrate community parameter (0.4), fi shery<br />

parameter (0.2), fi sh stocks parameter (0.3), plant parameter<br />

(0.1) and an optional conservation parameter (0.3).<br />

The major disadvantage was that <strong>the</strong> methodology was<br />

based solely upon Average Scores Per Taxon (ASPT)<br />

scores, a measure of <strong>the</strong> balance between pollutionsensitive<br />

and pollution-tolerant <strong>in</strong>vertebrates. This provides<br />

what is essentially an <strong>in</strong>dex of organic pollution<br />

without any accommodation of <strong>the</strong> overall faunal richness<br />

of samples, <strong>the</strong> fl ow velocity associations of species<br />

present, or <strong>the</strong>ir preference for ei<strong>the</strong>r perennial or ephemeral<br />

fl ow. The subjective manner of deduc<strong>in</strong>g severity<br />

<strong>in</strong>dex scores, based on targets that were produced on <strong>the</strong><br />

basis of adjustments to a start<strong>in</strong>g fi gure suited to sites<br />

considered to offer highest potential for ASPT, was also<br />

of major concern, s<strong>in</strong>ce it effectively skewed most lowland<br />

rivers towards low impact scores, despite <strong>the</strong> known severity<br />

of impacts with<strong>in</strong> chalk catchments at <strong>the</strong> time.<br />

Instream Flow Incremental Methodology<br />

(IFIM) and PHABSIM<br />

Instream Flow Incremental Methodology (IFIM) and<br />

PHABSIM have also been used to establish fl ow objectives<br />

from direct measurement (or modell<strong>in</strong>g) of resident<br />

biotic communities and <strong>the</strong>ir specifi c habitat requirements,<br />

although <strong>the</strong>se models have a tendency to be <strong>in</strong>accurate<br />

and <strong>in</strong>adequate, and at best require careful calibration to<br />

prevent distortion of results.<br />

The Surface Water Abstraction Licens<strong>in</strong>g Policy<br />

(SWALP) Methodology<br />

The SWALP methodology developed for <strong>the</strong> NRA by <strong>the</strong><br />

consultant eng<strong>in</strong>eers Sir William Halcrow & Partners Ltd<br />

<strong>in</strong> 1995 <strong>in</strong>corporates an environmental weight<strong>in</strong>g (EW)<br />

system, for which scores for sensitivity to abstraction are<br />

assigned to physical character, ecology and fi sheries characteristics<br />

(NRA, 1995). The catchment is fi rst assigned<br />

an EW score and thresholds are produced for a set of<br />

defi ned critical assessment po<strong>in</strong>ts (APs).<br />

The pr<strong>in</strong>ciple beh<strong>in</strong>d <strong>the</strong> SWALP ecological scor<strong>in</strong>g<br />

system was that <strong>the</strong> highest scores were allocated to<br />

species perceived to require coarse bed materials and<br />

rapid/fast current velocities, and lower scores were given<br />

to species which are common <strong>in</strong> still water habitats or are<br />

capable of withstand<strong>in</strong>g dessication. In contrast to <strong>the</strong><br />

SWK method, SWALP targeted environmental weight<strong>in</strong>g<br />

towards physical and biotic (ecology and fi sheries) attributes,<br />

exclud<strong>in</strong>g use-related amenity and recreational criteria,<br />

which were somewhat subjective. It also represented<br />

<strong>the</strong> fi rst attempt to br<strong>in</strong>g toge<strong>the</strong>r <strong>the</strong> perceived fl ow pref-<br />

erences of <strong>in</strong>dividual macrophyte and <strong>in</strong>vertebrate species<br />

<strong>in</strong> a biotic <strong>in</strong>dex concerned solely with fl ow dependency.<br />

However, <strong>the</strong> environmental weight<strong>in</strong>gs produced by <strong>the</strong><br />

SWALP methodology are ra<strong>the</strong>r crude and offer only a<br />

subjective basis for decision mak<strong>in</strong>g and a limited capacity<br />

to monitor improvement or determ<strong>in</strong>e <strong>the</strong> relative fl ow<br />

requirements with<strong>in</strong> or between river bas<strong>in</strong>s.<br />

The Lotic–Invertebrate Index for Flow Evaluation<br />

Index (LIFE)<br />

Follow<strong>in</strong>g <strong>the</strong> <strong>in</strong>vertebrate sensitivities to fl ow suggested<br />

by <strong>the</strong> SWALP methodology, Extence et al. (1999) developed<br />

LIFE. The method sought to l<strong>in</strong>k changes <strong>in</strong> river<strong>in</strong>e<br />

benthic macro<strong>in</strong>vertebrate assemblages to prevail<strong>in</strong>g fl ow<br />

regimes. The <strong>in</strong>dex is based upon <strong>the</strong> ‘primary fl ow associations’<br />

of different maco<strong>in</strong>vertebrate taxa at ei<strong>the</strong>r<br />

BMWP1 family or mixed-species level – deduced, for <strong>the</strong><br />

most part, from taxonomic keys. Calculation of <strong>the</strong> LIFE<br />

<strong>in</strong>dex score for a sample <strong>in</strong>volves <strong>in</strong>dividual fl ow scores<br />

(fs) for each scor<strong>in</strong>g taxon present <strong>in</strong> a sample obta<strong>in</strong>ed<br />

from a matrix. This matrix is based on <strong>the</strong> <strong>in</strong>frastructure<br />

of <strong>the</strong> biotic score system proposed by Chandler (1970)<br />

for assess<strong>in</strong>g water quality (Extence et al., 1999). Increas<strong>in</strong>g<br />

abundance of stand<strong>in</strong>g water or drought resistant<br />

species produces a lower fs score, whilst <strong>in</strong>creased numbers<br />

of <strong>in</strong>dividuals from runn<strong>in</strong>g water fl ow groups produces<br />

higher scores.<br />

Whilst changes <strong>in</strong> LIFE score may faithfully refl ect <strong>the</strong><br />

observed shifts <strong>in</strong> composition at sites undergo<strong>in</strong>g gross<br />

changes <strong>in</strong> habitat provision, <strong>the</strong> depletion of fauna caused<br />

by an earlier channel-dry<strong>in</strong>g event can produce spurious<br />

LIFE scores. This may occur, for example, by <strong>the</strong> temporal<br />

exclusion of molluscs normally associated with slack or<br />

slow-fl ow<strong>in</strong>g water or by rapid colonisation of blackfl ies<br />

(Simuliidae) and mayfl ies (e.g. Baetis spp) at a re-wetted<br />

site (Figure 8.4). These aspects of recolonisation usually<br />

produce a faunal composition with a higher proportion of<br />

velocity-dependent species and an artifi cially elevated<br />

LIFE score, even though taxon richness and diversity may<br />

rema<strong>in</strong> severely depleted. In <strong>the</strong> case of <strong>the</strong> <strong>River</strong> Misbourne<br />

(Appendix 8.3), LIFE scores were also found to<br />

vary signifi cantly between sites <strong>in</strong> close proximity (100 m<br />

apart) l<strong>in</strong>ked to spatial differences <strong>in</strong> habitat representa-<br />

1 The BMWP (Biological Monitor<strong>in</strong>g Work<strong>in</strong>g Party) scores<br />

refere to a system for assess<strong>in</strong>g water quality based on <strong>the</strong> macro<strong>in</strong>vertebrate<br />

assemblage. Each taxa is allocated a value from 1<br />

to 10 depend<strong>in</strong>g on its known tolerance to organic pollution. High<br />

scor<strong>in</strong>g taxa have lower pollution tolerance, and are thus an <strong>in</strong>dicator<br />

of good water quality.


<strong>Uncerta<strong>in</strong>ty</strong> Surround<strong>in</strong>g <strong>the</strong> Ecological Targets and Response of <strong>River</strong> and Stream <strong>Restoration</strong> 159<br />

tion <strong>in</strong> a relatively unmodifi ed chalk stream. Accord<strong>in</strong>gly,<br />

<strong>the</strong> LIFE score response largely depended upon <strong>the</strong> magnitude<br />

of temporal habitat change observed at an <strong>in</strong>dividual<br />

site. This highlights <strong>the</strong> need to select a representative<br />

variety of sites to assess <strong>the</strong> l<strong>in</strong>ear extent of environmental<br />

damage attributable to low fl ow.<br />

The MISINDEX<br />

An alternative fl ow dependency <strong>in</strong>dex, MISINDEX, was<br />

developed for use on <strong>the</strong> <strong>River</strong> Misbourne (Appendix 8.3)<br />

to accommodate cases where assemblages are transitional<br />

and depleted by previous dry<strong>in</strong>g. The MISINDEX works<br />

<strong>in</strong> a similar manner to <strong>the</strong> LIFE <strong>in</strong>dex, with <strong>the</strong> added<br />

advantage that it featured extended taxonomic coverage of<br />

aquatic species and <strong>the</strong> <strong>in</strong>clusion of water dependent (but<br />

not fully aquatic) species of Coleoptera. As with <strong>the</strong> LIFE<br />

<strong>in</strong>dex, <strong>the</strong> MISINDEX used fi ve fl ow-dependency groups<br />

and used abundance data for taxa <strong>in</strong> a similar manner to<br />

Extence et al. and Chandler’s matrices, but <strong>in</strong>troduced a<br />

category of water-related or marsh species that were not<br />

necessarily aquatic but depend upon wet places with<strong>in</strong> a<br />

river corridor, lake bas<strong>in</strong> or marsh. Moreover, ra<strong>the</strong>r than<br />

attempt<strong>in</strong>g to allocate <strong>in</strong>dividual species to a primary fl ow<br />

type association that was evidently subjective, <strong>the</strong> <strong>in</strong>tention<br />

was to defi ne <strong>the</strong>m by <strong>the</strong> limitations of <strong>the</strong>ir tolerance<br />

to fl ow cessation and to tell it like it is if a taxon was of<br />

low or unknown fi delity.<br />

Nei<strong>the</strong>r <strong>the</strong> LIFE or MISINDEX <strong>in</strong>dices explicitly take<br />

<strong>in</strong>to account <strong>the</strong> dispersal ability of species, or <strong>the</strong>ir ability<br />

to recolonise fl ow derogated sites after fl ow restoration, an<br />

important factor that requires fur<strong>the</strong>r <strong>in</strong>vestigation and<br />

could improve prediction of <strong>the</strong> likely time scale of<br />

impacts result<strong>in</strong>g from fl ow derogation or periodic channel<br />

dry<strong>in</strong>g.<br />

Detrended Correspondence Analysis (DECORANA)<br />

As a multivariate statistical technique DECANORA can<br />

be used to quantify similarities between samples or sites<br />

ei<strong>the</strong>r spatially or temporally. From such an analysis, it<br />

may be possible to produce mean<strong>in</strong>gful fl ow objectives<br />

determ<strong>in</strong>ed by observed empirical evidence of <strong>the</strong> l<strong>in</strong>ks<br />

between faunal composition or population sizes of <strong>in</strong>dividual<br />

species and river fl ows or o<strong>the</strong>r recorded environmental<br />

characteristics l<strong>in</strong>ked to hydrological variables <strong>in</strong><br />

a particular river reach. For this, longer term time-series<br />

of <strong>in</strong>formation are required than usually exist at present,<br />

as experience suggests that <strong>the</strong>re is considerable fl ux <strong>in</strong><br />

<strong>the</strong> composition of assemblages, <strong>the</strong> abundance of constituent<br />

species and <strong>in</strong>stream habitat characteristics l<strong>in</strong>ked<br />

to <strong>the</strong> fl ow history of sites and many o<strong>the</strong>r factors.<br />

APPENDIX 8.3 MONITORING THE<br />

ECOLOGICAL RESPONSE OF THE RECOVERY<br />

ENHANCEMENT OF THE RIVER MISBOURNE<br />

Introduction<br />

The <strong>River</strong> Misbourne, a 28 km long, ecologically valuable<br />

chalk stream situated <strong>in</strong> <strong>the</strong> <strong>River</strong> Thames catchment <strong>in</strong><br />

Berkshire, was believed to be amongst <strong>the</strong> rivers most<br />

affected by abstraction <strong>in</strong> <strong>the</strong> United K<strong>in</strong>gdom. Compounded<br />

by <strong>the</strong> effects of drought by 1996/1997, migration<br />

of <strong>the</strong> perennial head saw <strong>the</strong> river dry from its source<br />

to <strong>the</strong> middle reaches. A scheme to restore fl ow was implemented<br />

by <strong>the</strong> Environment Agency and Three Valleys<br />

Water, which comprised abstraction reduction and relocation<br />

of abstraction po<strong>in</strong>ts. A series of boreholes commissioned<br />

by <strong>the</strong> Environment Agency (EA) monitored <strong>the</strong><br />

hydrological response of <strong>the</strong>se measures, with water<br />

quality rout<strong>in</strong>ely monitored at four sites.<br />

The ecological response to <strong>the</strong> attempted fl ow recovery<br />

was monitored at six sites along <strong>the</strong> course of <strong>the</strong> river,<br />

<strong>in</strong>clud<strong>in</strong>g sites that had dried down and those that had<br />

reta<strong>in</strong>ed fl ow. Sites with differ<strong>in</strong>g characteristics were<br />

chosen to assess response more thoroughly, although <strong>the</strong><br />

variation between sites meant that differences between<br />

sites needed to be compared. For example, Sites 1 and 2<br />

were dry <strong>in</strong> 1998; Site 3 rema<strong>in</strong>ed dry until 2001, reta<strong>in</strong>ed<br />

fl ow until spr<strong>in</strong>g 2003, but had dried down aga<strong>in</strong> by<br />

autumn 2003; Site 4 was dry only <strong>in</strong> autumn 1997; and<br />

Sites 5 and 6 reta<strong>in</strong>ed fl ow throughout.<br />

An attempt was made to monitor ecological response<br />

us<strong>in</strong>g a series of standard survey methodologies typically<br />

applied twice annually (often spr<strong>in</strong>g and autumn) over a<br />

three-year period (1996 to 1999). Changes <strong>in</strong> vegetation<br />

with<strong>in</strong> <strong>the</strong> corridor was monitored with Phase 1 and 2<br />

Habitat Survey coupled with <strong>River</strong> Corridor Survey (RCS)<br />

and <strong>in</strong> <strong>the</strong> channel by Macrophyte Survey (MS) accord<strong>in</strong>g<br />

to changes <strong>in</strong> habitat conditions shown by <strong>River</strong> Habitat<br />

Survey (RHS). Birds were monitored us<strong>in</strong>g Common Bird<br />

Census (CBC) and W<strong>in</strong>ter Atlas (WA) surveys. For<br />

mammals, bat transect surveys were conducted and <strong>the</strong><br />

presence of otter Lutra lutra and Water voles Arvicola<br />

terrestris was to be recorded dur<strong>in</strong>g RCS.<br />

Macro<strong>in</strong>vertebrates and fi sh were monitored with more<br />

specifi cally designed sampl<strong>in</strong>g regimes. In <strong>the</strong> case of fi sh<br />

this provided an opportunity to compare two monitor<strong>in</strong>g<br />

techniques: standard depletion electric fi sh<strong>in</strong>g with<strong>in</strong> stopnets,<br />

and po<strong>in</strong>t-abundance sampl<strong>in</strong>g by electric fi sh<strong>in</strong>g<br />

(PASE) (Copp and Peñáz, 1988; Perrow et al., 1996). Follow<strong>in</strong>g<br />

this, <strong>the</strong> longer term response over a fur<strong>the</strong>r fi ve<br />

years was undertaken us<strong>in</strong>g PASE.<br />

The project had a general aim to restore a native fi sh<br />

community of Brown trout Salmo trutta and bullhead


160 <strong>River</strong> <strong>Restoration</strong>: <strong>Manag<strong>in</strong>g</strong> <strong>the</strong> <strong>Uncerta<strong>in</strong>ty</strong> <strong>in</strong> Restor<strong>in</strong>g Physical Habitat<br />

Cottus gobio, but no specifi c targets were set; not only<br />

was <strong>in</strong>formation on <strong>the</strong> fi sh population scarce, <strong>the</strong>re was<br />

also no move to ga<strong>the</strong>r more. However, <strong>the</strong> connection of<br />

<strong>the</strong> Misbourne to <strong>the</strong> <strong>River</strong> Colne was thought likely to<br />

provide a source of colonists at least to <strong>the</strong> lower reaches.<br />

Shardloes Lake, with a direct connection to <strong>the</strong> river, was<br />

also thought likely to operate as a refuge for fi sh <strong>in</strong> <strong>the</strong><br />

upper reaches, although this was also thought to have dried<br />

out <strong>in</strong> 1997/98.<br />

Comparison Between Fish Survey Methods<br />

Each site was sampled twice annually (spr<strong>in</strong>g and autumn)<br />

between July 1996 and October 1999 us<strong>in</strong>g both depletion<br />

fi sh<strong>in</strong>g and PASE. The same electrofi sh<strong>in</strong>g equipment was<br />

used for both methods. Depletion electric fi sh<strong>in</strong>g was<br />

undertaken with<strong>in</strong> a 100 m section created by <strong>the</strong> use of<br />

stop-nets at ei<strong>the</strong>r end to prevent <strong>the</strong> escape of fi sh. Two<br />

operators, each with a s<strong>in</strong>gle anode attached to a separate<br />

100 m cable and a hand net fi shed <strong>the</strong> section, explor<strong>in</strong>g<br />

<strong>the</strong> entire area with sweep<strong>in</strong>g movements. Two runs were<br />

conducted at all sites, apart from at Site 6 <strong>in</strong> spr<strong>in</strong>g and<br />

autumn <strong>in</strong> 1997, where emergent vegetation seriously<br />

hampered fi sh<strong>in</strong>g and only one run was attempted. All fi sh<br />

captured were reta<strong>in</strong>ed <strong>in</strong> one or two b<strong>in</strong>s carried by additional<br />

personnel and processed (identifi ed and measured,<br />

with some fi sh also be<strong>in</strong>g weighed to calculate length–<br />

weight regressions) at <strong>the</strong> end of each run, before release<br />

downstream of <strong>the</strong> stop-netted area. Us<strong>in</strong>g <strong>the</strong> numbers of<br />

each species captured <strong>in</strong> each run, an estimate of <strong>the</strong> total<br />

population size of each species was generated us<strong>in</strong>g <strong>the</strong><br />

maximum weighted likelihood model of Carle and Strub<br />

<strong>in</strong>d. m -2<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

(1978). Us<strong>in</strong>g <strong>the</strong> area of <strong>the</strong> section sampled (100 m ×<br />

width of <strong>the</strong> channel), quantitative estimates of numerical<br />

(<strong>in</strong>d. m −2 ) and biomass (g m −2 ) density were calculated.<br />

Prior to PASE, a volt meter was used to calibrate <strong>the</strong><br />

gear to provide a voltage gradient of 0.12 V at 45 cm away<br />

from <strong>the</strong> 40 cm anode. Such a voltage corresponds to <strong>the</strong><br />

m<strong>in</strong>imum effective voltage at which <strong>in</strong>hibited swimm<strong>in</strong>g<br />

occurs (Copp and Peñáz, 1988). The effective sampl<strong>in</strong>g<br />

area was thus calculated to be 1.3 m 2 from which quantitative<br />

estimates of abundance and biomass were calculated.<br />

A total of 50 po<strong>in</strong>ts were sampled with<strong>in</strong> a 200 m section<br />

of river directly upstream of <strong>the</strong> stop-netted section at each<br />

site. As <strong>in</strong> <strong>the</strong> depletion fi sh<strong>in</strong>g, <strong>the</strong> anode was attached to<br />

a 100 m extension cable from <strong>the</strong> control box, which was<br />

sited mid-way along <strong>the</strong> section. Po<strong>in</strong>ts were taken <strong>in</strong> a<br />

stratifi ed random manner at 4 m <strong>in</strong>tervals, with <strong>the</strong> electric<br />

fi sh<strong>in</strong>g operator mov<strong>in</strong>g diagonally upstream from bank<br />

to bank, and sampl<strong>in</strong>g a pre-determ<strong>in</strong>ed number of po<strong>in</strong>ts<br />

<strong>in</strong> <strong>the</strong> littoral marg<strong>in</strong>, which was determ<strong>in</strong>ed from <strong>the</strong> relative<br />

width of <strong>the</strong> marg<strong>in</strong> to that of <strong>the</strong> channel.<br />

Overall, both techniques tended to sample a similar<br />

number of species (Wilcoxon signed ranks test, n = 17,<br />

Z = −2.84, p = ns) and produce similar estimates of total<br />

biomass (n = 17, Z = −1.87, p = ns), but signifi cantly different<br />

estimates of total abundance (n = 17, Z = −3.10,<br />

p < 0.01) (Figure 8.5). PASE produced <strong>the</strong> higher estimates<br />

of total abundance on account of its tendency to<br />

produce signifi cantly higher estimates for bullhead (n =<br />

13, Z = −2.69, p < 0.01) and Ten-sp<strong>in</strong>ed stickleback Pungitius<br />

pungitius (n = 9, Z = −2.67, p < 0.01). There was<br />

little consequence of <strong>the</strong> production of signifi cantly higher<br />

biomass estimates for both <strong>the</strong>se species (p = 0.01 and<br />

gm -2<br />

Spr<strong>in</strong>g Autumn Spr<strong>in</strong>g Autumn Spr<strong>in</strong>g Autumn<br />

Spr<strong>in</strong>g Autumn Spr<strong>in</strong>g Autumn Spr<strong>in</strong>g Autumn<br />

1997 1998 1999<br />

1997 1998 1999<br />

PASE<br />

8<br />

6<br />

4<br />

2<br />

0<br />

DEPLETION<br />

Figure 8.5 Comparison of numerical (<strong>in</strong>d. m −2 ) and biomass (g m −2 ) density estimates obta<strong>in</strong>ed by PASE (mean ± 1SE shown) and<br />

depletion fi sh<strong>in</strong>g (density derived from total population estimate shown) over time at Site 5, dom<strong>in</strong>ated by small species such as<br />

sticklebacks, stone loach and bullheads

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!