2A490 ae UNIVERSITY OF 4 a Been ALTERRA (a Y ¢ ‘4 CAMBRIDGE a f . enpitenmental pak WAGENINGEN NL ?) : UNEP WCMC REVIEW ON THE ECONOMICS OF BIODIVERSITY LOSS: SCOPING THE SCIENCE ENV/070307/2007/486089/ET U/B2 FINAL REPORT (draft version as of 19 May 2008) Balmford, A., Rodrigues, A. (University of Cambridge), Matt Walpole (WCMC), ten Brink, P., Kettunen, M. (IEEP) & Braat, L. and de Groot, R. (Alterra) May 2008 This report has been produced for the European Commission under Contract ENV/070307/2007/486089/ETU/B2. Scientific lead: Andrew Balmford and Ana S.L. Rodrigues Conservation Science Group, Department of Zoology, University of Cambridge, Cambridge, United Kingdom Partners: Patrick ten Brink and Marianne Kettunen IEEP — Institute of European Environmental Policy, Brussels, Belgium Matt Walpole UNEP — World Conservation Monitoring Centre, Cambridge, United Kingdom Leon Braat and Rudolf de Groot Alterra, Wageningen University and Research Centre, Wageningen, The Netherlands External experts: The expertes that provided direct input into this project are listed under each theme. Acknowledgements: To all the experts that contributed as authors, reviewers, contributors or otherwise provided information useful to the thematic reviews. Also to the interns that worked on each task (see participant list in each theme) To Ben Phalan and Bernardo Strassburg for helping with case-studies for the glossy report. To the members of the Zoology Department that created the logistical conditions for this project, in particular Julian Jacobs, Nanna, Neil, David Clarke, and Keith Griffiths. TABLE OF CONTENTS 1 DNPRODUG TION: sesctsrcccesscscsssaactcescccsucescsucn otaceseonendze toca ete nettteans noel ceateuauscescesuncesncsecsesessese 1 1.1 RACK ON OWING Ei cccescacesccsssesaccncescaceauconscerscesnconstermectecctentes tenaccnate tc tesceacecnateseceneraatreerence 1 1.2 This Scoping the Science project..............csccsssscescessesecssssssessssseesssseeseeseessesseesees 1 1.3 AMSG PO Hit seescecteenccess cece sei she cavatsenssueeceesacsctecsseeeterseetectaddtatstascheossussceseccacesscaecvecyoes 2 2 A CONCEPTUAL FRAMEWORK FOR THE REVIEW......cscssssssssssssssssssssssssscsssssssseesaseseesees 4 2.1 Why we need a conceptual framework.................cssssscsccsssssccssceceeseeesseeseeeseseees 4 2.2 Initial considerations and assuMptions,.................scsccsssscsseseesssscecsscesssssssesseseesess 4 2.2.1 Waltin Wd NOUUre nc .ccseccosceccencenconcsesscceseaceseezsencvessessontcccenescecesesesussastecrsasecse 4 Dele, SCOPE Of ING REVIEW scccccsessescesessssseresccstsccesosceasscessecresvatsuoestsreccesctecctsatsssttecesees 6 2.2.3 Valuuing marginal ChANGe...........ccecsecsssesssceceecsescssscescsescsesssecsscssscssessecsonceees / 2.2.4 The need to be spatially explicit ......c.sssssssessesscsessesscssessesesscssccsescesessesseeseeeees 8 2.3 The conceptual framework..............csscssccsccessessssssssssssecsescecesesessessesssesseesecseeseesees 8 2.3.1 Defining appropriate policy actions based on the drivers of loss............... 10 Deed, Defining the counterfactual states Of the WOT ...........ssssseeeeseseseeseessoees 11 2.3.3 Quantifying and mapping how the biophysical provision of benefits is Affected by the policy ActiON ...........sesscsecseccesesccsccscesccsesceccesescsscescsecssesssecsecsecscesocseces 12 2.3.4 Quantifying and mapping the economic value of changes in benefits derived from the Policy ACtiON..........s..cseoceesceesseesssessesssessseccsesssecssnscscssecesessscssecssnscenss 16 2.3.5 Quantifying and mapping the overall economic value of changes in benefits derived from the Policy ACtON .......s.ssscseeseeseersessesssssesssescnsesscessesseessesseseasees 21 2.3.6 Quantifying and mapping the costs of Policy ACTION .....s.sssssseesseesereesereees 22 Deaf, Quantifying and mapping the net economic consequences of policy action 23 2.3.8 Evaluating the adequacy/desirability of the policy ACtiON.L.........1.00.00.00000 24 2.4 Summary of key points from the conceptual framework ..................sssseeees 24 2.5 PAN CI CLP ANUS eesecesscecsescuecosecocescerseetssesctsuccccastessscadseacscnccaceccercacesarsstocascuccssttencscasses 25 3. A STRATEGY FOR REVIEWING KNOWLEDGE ON THE LINKS BETWEEN WILD INATURE/ANDIFIUMAIN WELUBBING ceccorecccstncsscesesccstsesecssesroescssssesecsesesecttecscossessecsosaccescacceesee Di 3.1 Mhe nee dilo rials trate sy,ccccccce er ececceetcssoaneenacteacseacccecerastaasuassacesarmtscateatsanceesacccere 27 3.2 The Millennium Ecosystem Assessment as a starting POint .............cscseseeees 27 3.3 The need to avoid double COUNTING ...............s.ssssccssssssscsssessssssssesssccssssssseaseaesrees 28 3.4 Proposed classification of ecosystem processes and benefits ..............s.s0s00000 29 3.5 _ Partitioning the links between wild nature and human wellbeing as a basis for the thematic reviewS...............ccsccssccscessssscsscsscsssssssscssssssscsssssssessessesseeseeseeseeeees 35 Op INCI TIC REVIEW Slecsccerscccecceccesessseetnaccaccessasscarsacccasncvessocesdeensacaseaenactacsatsassacsansaces 38 3.7 _ Relationship between the proposed classification and the Millennium Ecosystem Assessment framework..............sccsscssscssssscsscsscescescesssscsscssesessnsesssasensesseences 40 3.8 Relationship between the proposed classification and the Total Economic Value framework ..............ssccssccssccssessscssssssssssesscessescessesessncsenssesecesseesscensens 42 Seeman Ee ALGUICH PD ATLAS fee secescsnesacsatesencestaansasacveccdcnesaaransassasactassascerssechonvanescsaacesessessscessssaemnsea 45 iil 4 THEMATIC REVIEWS. ......sscssssssssssssesssssesecensesssssestesssssescessessesssscssnssscsacensecsecsensoessessaseass 46 4.1 Wild crop pollimation...............scssccsscsssssssssscssscssssssssssssscssssessseesesseessesasenssenseessens 48 4.1.1 Why is this process important for human wellbeing? .....s.scsesereererereees 48 4.1.2 What are the overall trends in the provision of wild crop pollination?..... 48 4.1.3 How is the provision of wild crop pollination affected by changes in WEIN ALUT.C Oicncsesconsscsoncssaaccaccacscesscsonstotastostssodcetasetesestacsctaascocssbassssseceseccsnncessonssacnssanneta= 49 4.1.4 What are the main threats to the provisioning of wild crop POLLINATION? ......cceccerrsscsessssccsscesccccsccescccnsccsscsaccnscscescssscseacsoncesscenseseassasensccecsscccnneenaeenees 52 4.1.5 Are abrupt changes likely in the provision of wild crop pollination?........ 52 4.1.6 Can we quantify and map the global provision of this wild crop Pollination and how it Might ChANGE? ...ccscsccsscsersceseeeereserseesesserssssnsnseeesscensensesaees 53 4.1.7 Insights for CcOnOMicC VALUATION.......c.sccssccsssssseereerseesecesseescenssensssseeesesnseenees 57 4.1.8 SOMME HEVIMESOUTCES scestesscrsesrereeqcotescactaeceesecestsccs ranccstoeceocssececoccoscsscssancezesansees 57 4.1.9 IRATTICIDANISecresncenscstectrectrtrretetscatsacssecsscecuscesteasesecscccuccsnscnactscsautreassseseccees 58 4.2 Biological control im Crops ...........ccsccsssssessessessessessessessesnseccsccsncesessessessesscssesacenees 59 4.2.1 Why is biological control in crops important for human wellbeing?........ 59 4.2.2 What are the overall trends in the biological control Of CropS? .....100.100006 60 4.2.3 How is the provision of biological control in crops affected by Changes in wild nature? ...........cccsccsssccsscssscsssssssecssesccessssscssssessssesscssasssssssscacssnsseesees 61 4.2.4 What are the main threats to the provision of biological control in crops? 63 4.2.5 Are abrupt changes likely in the provision of this benefit/process?........... 63 4.2.6 Can we quantify and map the global provision of biological control in Crops, ANA how it Might CHANGE? ..esceccssserssrssreereeeersecersersecsersesssssesecsesesessseseceeeeees 64 4.2.7 Insights for CCONOMIC VALUATION.......cscesssssceseesscsssessecescessssssesssessssssenseessees 65 4.2.8 SOME) EVI CSOUTCES ccccesecccssaccsecatsnccsccsessossecuctsezaccacacccesoosescsecossescassazsesecenseane 65 4.2.9 IROTTUCIPONISiecsnscsnsecocastacvoesacssctscceccestancecessecsesecsaccaccseresssissasaasaccaccaesceceucsaeeeees 65 4.3 Genetic diversity of crops and livestock ..............sccscssssssessscsssssssscsssssssssseseseees 66 4.3.1 Why is the genetic diversity of crops and livestock important for WUMmanbw CHB CIN Ge rrrccsectcosttsesncchscoctsestenccctenceasoneteceacerenassucecesceeceesnastastesasececsesacevescarsaes 66 4.3.2 What are the overall trends in the genetic diversity of crops and LEGO 5% cecececgs2c802000000- 00005 2600 CCE CEO DOC ECE ESOC OCRO CB CoS Sco GHOCHACCOCGSSCoSC ES EEA AE eeepc coocIacEC 67 4.3.3 How is the genetic diversity of crops and livestock affected by changes EM WIE NOQLUTE? vccccecccsscesuscccacsotesessasacsecestanectases tavetoesettocseaasseziesdensssuecsid ccsdestoesesocttvotseosee 67 4.3.4 What are the main threats to the genetic diversity of crops and UL VESLO CK Zasvensanssntesscstnices sssesstaanscszsateateascecenatatasneaneaseateacancatucecessarcarcarnatctertensserteceetecees 68 4.3.5 Are abrupt changes likely in the genetic diversity of crops and VEVESLO CK idperecescnccnascentenrentsnteatonesasccasesssassareasedcesceacsstcatccsanecanscceerrateteecsccseieeseoc seer one ete 68 4.3.6 Can we quantify and map the global diversity of crops and livestock, ANGHOWND MICH CHANGE essere sserscancecesncescencertcccereeceertectersectestsassscereecceesceceeoteteotnieee 69 4.3.7 Insights for CcOnOniic VALUATION .....1.scsccsssseseeresssssesersessessessasessasessscessacsnesses 70 4.3.8 SOME: ROY PESOUTCES i ocsxccacsacoccsecsurtccuucsuceactertsuctnevnssaston toast iaesesiivickatot eee atest 70 4.3.9 IP AVEICIPOMIS pandetantenteacntcccnssnenessarateiaes caraacuasnenseeasnes eres rereeceteer tee eeeeee eee 71 4.4 Soil quality for crop production. ..............sscscscscscsccesssecesssesesesesesesescsesesesesessscees 72 4.4.1 Can we quantify and map the contribution of wild nature for the provision of soil quality for crops, and how it might Change? ....s.c.s.sssesseseeserereeeseeess 76 4.4.2 SOME/K CY FCSOUTCOS «0.5 ccessicscsucsccscsersasccte Rea Ae sce ess 76 4.4.3 ROT UICID GNU S\accnsceszscssessoncsssectescesterecsmteencveccterstescnceree teeter nese eeeteeseeeas Tif, 4.5 Wives CKaetacsssssncarccespnesasesessscceactoescotestastesseusseaeatesnaes saceneceecae site esacesseteteate reese sce 78 4.5.1 Why is the production of livestock important for human wellbeing?........ 78 4.5.2 What are the overall trends in the production of livestock? ........0..0000000006. 78 4.5.3 How is the production of livestock affected by changes in wild nature?... 78 4.5.4 What are the main threats to livestock production in natural BTASSTANAS Tee Mines sate seeecon tee sce ecn acres tac s teeta cece aa ecserse cto nace 79 4.5.5 Are abrupt changes likely in the provision of livestock in natural QT OSSIANAS PrevcccsancsvscssassesstocsunscsscstoCucsucveewesaseussicatestscsccereate eee teeter ete ee ak eakeses 79 4.5.6 Can we quantify and map the contribution of wild nature for livestock Production, and how it might CHANGE? .....ssscesesesseerssssssssssecsssssssecessescssssssessssssessesees 79 4.5.7 Insights for CCONOMIC VALUATIONL.......scscesercesssesssscescessesseessesesseessesssssensssees 80 4.5.8 SOME KEY TES OUNCOS cc tzcsccescosccetsosocsorsusteiasteesiesaduseteeoeiotes eveate acaseaethetasiatse 80 4.5.9 PATULGID ANUS s28. Sosssece tt ce cectac acest sev eeaeeec ttn eeteantet eae seeaco ee ceatcasotent tier cas 81 4.6 aus Marine fish Cries). sccscssc.cdesccoeeccors 96 4.7.3 How is the provision of inland fisheries and aquaculture affected by GCHANGESiIN Wild MAtUTe 2 sccccectcesettentnocscctcescoaseateasesscccecesacctevcesietectessvetcacsscceseustoctaasaaseess 96 4.7.4 What are the main threats to the provision of inland fisheries and CUTTER Ree ECP PEED TTT CCRE EE EERE CECE EEE PEE EEE CECE CEP RPE EEE 99 4.7.5 Are abrupt changes likely in the provisioning of freshwater fisheries ACL OU OCHILUNG Aenvcnsenetnesusccacsuoctastuascastessaseasesennasuastatanvenceecactasepeucee:ceaeecettctetererseeeee 100 4.7.6 Can we quantify and map the global provision of inland fisheries and aquaculture, and how it might ChANGe? ......1.seeeesssssessssessessesssssssssrsesssssenscssenscesees 100 4.7.7 Insights for CCONOMIC VALUATION... seccscesssessescesssssscesescessesssesssscassscssensesees 102 4.7.8 SOME/KEY TOSOULCES ss.cc.acssecee eae eater eh von bad sauhace eee obatnanesbbuctiebiestsstes 103 4.7.9 PQTUCIP ONES ii re hase ES ccs 104 4.8 Wild! animal products ssc.ccccccccseccceressersesester eset teeter reo aa toon seebaccencteasotessesees 105 4.8.1 Why is this benefit important for human wellbeing? ....c.cccscssssseeerereeeeees 105 4.8.2 What are the overall trends in the provision of wild animal products?... 107 4.8.3 How is the provisioning of wild animal products affected by changes PAWID NOLUTE 2 mircsessckacsnsoscnsosdencncerssseeetee ee tt RT eee eens 108 4.8.4 What are the main threats to the provisioning of wild animal POTTS Se ene rectpoer eco cco coc ICL ASS SCG GI o2C0030 0005220 Ne GOCE CORO CECEEEOS ECL PECECODCHCCECEB ECO PEESSS 113 4.8.5 Are abrupt changes likely in the provision of wild animal products?..... 114 4.8.6 Can we quantify and map the global provision of this benefit/process ANGMOW TMB ht CRANGOL rerrecs-ccecscecensnereternesereenenctoteceenetersten er aseaceee tenet sas sacae staat easeee 115 4.8.7 Insights for CCONOmMIC VALUATION. ....ccccceccecessersersessesseessessessessessessessessessenses 119 4.8.8 SOME MEY TESOUTCES .: savesccsacsssestterrsrcencottaeteesed cena taceceea ee soeeeie cote tedetasenssts 120 4.8.9 DPATLICIPORIS oo coe cree ccasstete tee C Tee hii aca gasatsiee 120 4.9 Fresh water provision, regulation, and purification ..............ccsccccsssesesceeeees 122 4.9.1 Why is this process important for human wellbeing? .........s0ssseeeeeeeee 122 4.9.2 What are the overall trends in fresh water provision, regulation, and DATTA LGU (teres ec OR ECE EE ECE CECECEE SE CPPCC ATCC PELE OPE EEC CED 124 4.9.3 How is the provision, regulation, and purification of fresh water Affected by changes in Wild NAture?.....sssssceecessseecsssesssseccssccsecsssscsececsssessessesseessaces 124 4.9.4 What are the main threats to the provision, regulation and DUTU{TCATONVOSA TES W ALCT, ooccecscencsucuccocvccsccotvcterectacectacecsvsesescsacssssasestsisssessisetsssacereveee 129 4. 4. 4. 4.9.5 Are abrupt changes likely in the provision and regulation of fresh water? 130 4.9.6 Can we quantify and map the global provision, regulation, and purification of freshwater, and how it might ChANge?.......scecscesscsessssssssssssessescenes 131 4.9.7 Insights for economic ValUation.......s.scsecescsessesssesscesscenssessssssecseacesscenceeses 134 4.9.8 SOME /KEY PESOUTCES -saceeccectszesccsecsusvasvasseesoscsetncssenesaceuuccscarestcssssseasasisaseessets 134 4.9.9 IROTULCIDANUSjesressencutesceasetcstescunccsenceccuscetcusteacetneeeraarnecasnacnrsecorsasesescurtoceevererts 135 NO Wialdliiplamtefibresiccs.ccscscrece-cuesscectcceasescuescsssecscesseactasoncotsnscssencorsasocscsersasetacsisscrs 136 MA Wildinvedicunallplamtsscsscc.crscecccetearesctensecesssestsnectoseresessactessasarsacessnccevescesssataece 137 12 Outdoors activities related to Mature .............scscsccscsssessctscssetscsssseeeseeeseeceaces 138 4.12.1 Why is this benefit important for human wellbeing? .......0..ssseseeeeseeees 138 4.12.2. What are the overall trends in the provision of this benefit?..............00 139 4.12.3 How is the provision of this benefit/process affected by changes in WELD OLUTC Yownsccucecanectncer sesso een cotene ees senenteotesnectcnsnasecatnetursancatcesertacecsacestucassateseetats 140 4.12.4. Whatare the main threats the provision of this benefit? ...........s1sees000 143 4.12.5 Are abrupt changes likely in the provision of this benefit? ........1...0.1000 144 4.12.6 Can we quantify and map the global provision of this benefit and how WL UGH CHANGED reccecceerssreeeccrsctst etter taatnseenteaneeneasucctescetanssesaecsesteescesuesccsncccsccsceesossesss 144 4.12.7 Insights for economic VAlUation.....s.sccscereereeresrerreerserscsrserserensenseesensseseesees 146 Vi ANU ZESMS OMEIK CITES OUT CES -oxn.cczactcaseacersacsusessactasccarceesscocsesncuansuasvocsesaceascucceaesaaeseess 147 4.12.9 Participants in this SUb-review..........scssccsscesreerecesscssseesscesseesserseesesseesseess 147 4.13 Regulation of natural hazards .............scsscsscscscsssecsssssessscesssseessessesesesseeaeesees 149 4.14 One-off use benefits ...............cescsscecessscescsscsscsscsscsssersesssessscsessssseaseasessensensens 150 4.15 Nom-use bene fits.............cscccccsscsssssssssccsssscsscssscsscsssssssesessessscssessaeseesseasensseneeess 151 Global climate regulation. .................:ccsscssssssssssccesscsscssscscssessecsssssscsscsssessseesenssesssenseeass 152 4.15.1 | Why is global climate regulation important for human wellbeing?........ 152 4.15.2 How is global climate regulation affected by changes in wild nature?... 152 4.15.3 Can we quantify and map the value of wild nature for global climate regulation, and how it might CHANGE? .....scsssscsreesessecseerscrscssceceeseessecsecsecsecseneesseescees 153 ASU SYA WMP AVLECLP AN Stcsrsctsscccsosssesersucsctsnesecsercentsnstencencseeenesonccaceestassacenssatsansansseseesoers 154 4.16 Unknown benefits or processes ...........s.ssssscsscesssesscsssssssscescsessessesessssesssesseesees 155 ANP S COM ANOS seetescentsessssarcasoatescecsnaccnccesaststecnactancsunseacsuatesarsnacsncecaecsuresatsustspessestaatsness 156 A SUSI BOT LL CLD ANUS) :cncsssscsscenaesasseseeareatsascarenttatseacsaneeasnavauscatesessescecacesttaceacasasasseaanees 162 4.18 Prioritisation of recommendations for future WOrK.............scsscccseessessessseeees 163 ANUS ST OSTLCID AMS |ecencescsasssaacsansseassenascatsonscarecssescvaconseccetassuccosatsnceuasteatscareatsectsnctes 166 5 INVENTORY OF RESEARCH ORGANIZATIONS, PROGRAMS AND RECENT LITERATURE DEALING WITH ECONOMICS OF BIODIVERSITY LOSS ........scsscsecsesssssecseeeees 167 5.1 Write rate TO VICW cecessccessecesecceccececerectoresteresscscescasvscecsecssaccnscceeseorececesecuecsccesesees 167 5.2 Review of organisations, programmes and NetWOrks............ccscessesesresesers 167 GPIRERBRENGES esresecteecsseseescrccssssttacvorconcenseactaccacesssstacnncacccessescestessacenstrarnercaccnsesracemataasn= 170 Annex 1 — Cost and benefits of agriculture CONVErSION............ssscsscesecessesssssseseesens 207 Annex 2 — Framework development and consultation. ............scsssssssssssessesesseesees 212 Annex 3 - Practical strategy followed in the thematic reviewS............cscssecseseseeers 213 Annex 4 - Literature review “Economics of Biodiversity Loss” ...............ssssssesees 216 Annex 5 - Inventory of research organizations, programs and projects dealing with “Economics of Biodiversity LOSS” ..............ssssccssscsrsssssssssesssssscesscsessnsesacsesseseens 233 Annex 6 - Key Internet Resources on Biodiversity and Ecosystem Services .......... 241 Vil 1 INTRODUCTION 1.1 Background The recently released Stern Review on the Economics of Climate Change (Stern 2007) concluded that it makes economic sense to invest in mitigating greenhouse emissions to avoid the worst effects of climate change rather than face the consequences of failing to do so. This proved highly influential in persuading politicians and economists of the need to act on climate change now. The G8 decided in March 2007 to initiate an analogous “Review on the economics of biodiversity loss”, in the so called Potsdam Initiative: 'In a global study we will initiate the process of analysing the global economic benefit of biological diversity, the costs of the loss of biodiversity and the failure to take protective measures versus the costs of effective conservation.'. This proposal was endorsed by G8+5 leaders at the Heiligendamm Summit 6-8 June 2007.” This study consists of two phases: - Phase | is a preparatory stage including several studies to be ready for the CBD COP9 (Conference of the Parties to the Convention on Biological Diversity) in May. - Phase 2 is the consolidation stage that will produce the full Review, to be ready in October 2009. The study is being supported by the European Commission (together with the European Environmental Agency and in cooperation with the German Government), notably through contracted studies such as the present one. Leading economist Pavan Sukhdev as been appointed as an independent Review leader for Phase 2. 1.2. This Scoping the Science project The “Review on the economics of biodiversity loss: scoping the science” project (henceforth, the Scoping the Science project) is one of several Phase 1 projects, running from 14/12/2007 to 30/4/2008. According to the invitation to tender: “The objective of the current study is to provide a coherent overview of existing scientific knowledge upon which to base the economics of the Review, and to propose a coherent global programme of scientific work, both for Phase 2 (consolidation) and to enable more robust future iterations of the Review beyond 2010.” The specific tasks to be carried out in the context of the project are: 1. Elaborating the conceptual framework for the Review for biodiversity; 2. Establishing a working relationships with research networks and programmes; 3. Reviewing existing ecological knowledge needed to carry out the Review: 4. Identifying critical gaps in knowledge; 5. Presenting key findings; 6. Elaborating a proposal for a 1-2 year programme of work to provide case studies; 7. Elaborating a proposal for a 2-3 year programme of work to fill critical gaps; 8. Building an inventory of relevant research networks, programmes and projects; and 9. Providing an inventory of internet resources. The outputs of this Scoping the Science project will provide background and recommendations for Phase 2, when the approach to be followed in the Review will be defined in detail. 1.3 This report This report presents the final results and recommendations from the Scoping the Science project. The Conceptual Framework for the Review on the Economics of Biodiversity Loss (Task 1) is detailed in Section 2. Note that this has a significantly broader scope than the remaining tasks in this project. The conceptual framework was developed for the whole Review, covering the fuii integration of ecological and economic sciences. This task was led by the University of Cambridge. The bulk of this report covers the most substantial task of the project, of providing a coherent overview of existing scientific knowledge (Task 3). Given the specific mandate that such review should provide an adequate basis for the economic valuation, we ensured that the definition of the review themes selected was appropriately defined for this aim. Section 3 presents the conceptual background and justification for the selection of the set of themes that form the basis of the review of the ecological knowledge. Section 4 presents the results of the thematic reviews. Gaps in knowledge (Task 4) and key findings (Task 5) are discussed in the report for each theme. An extensive network of experts was created in this review (Task 2; see Error! Reference source not found.). This work was led by the University of Cambridge, with the support of UNEP-WCMC. A synthesis of the recommended priorities for Phase 2 (Task 6) as well as in the longer term (Task 7) is presented in section 4.18. This task was led by the University of Cambridge, with the support of UNEP-WCMC. The project also built an inventory of relevant research networks, programmes and projects (Task 8), with the aim of expanding the relationships with these in the future (Task 2). To support this exercise, a review of the recent literature (i.e. publications post the 2005 Millennium Ecosystem Assessment) dealing with economics of biodiversity loss was conducted (Section 5; with the list of identified publications and networks presented in Annex 4 and in Annex 5). This work was led by Alterra and the Wageningen University. Finally, the project also provides a review of key internet resources on biodiversity and ecosystem services (Task 9). This task was lead by UNEP-WCMC and its outcomes can be found in Annex 6. This review was specifically focused on web-available material related to biodiversity research and ecology (i.e. as oppose to biodiversity economics). Thus, its outcomes largely complement the information on research initiatives / networks and literature on economics of biodiversity loss above. 2 ACONCEPTUAL FRAMEWORK FOR THE REVIEW 2.1 Why we need a conceptual framework The scope of the Review on the Economics of Biodiversity Loss is extremely ambitious: analysing the global economic benefit of biological diversity, the costs of the loss of biodiversity and the failure to take protective measures versus the costs of effective conservation (Potsdam Initiative). The scope of the Review is also very wide, as the benefits that we obtain from biodiversity are extremely varied. Tackling this immense task will require building from the knowledge and the skills of an extensive network of research teams with very different expertises and backgrounds. The conceptual framework is a proposed operational roadmap for Phase 2 of the Review on the Economics of Biodiversity Loss. It presents a strategy for partitioning the Review into well-defined and self-contained pieces that can be tackled by teams with the appropriate expertise. It then presents the method for linking these pieces together in order to provide a reliable answer to the Postsdam challenge. The framework is particularly important for ensuring a smooth link between the ecological and the economic aspects of the Review. The Review will need to face substantial gaps in knowledge and so the results will inevitably be plagued by uncertainty. Dividing the Review into modules allows for the identification of the areas where the main gaps are. Understanding the links between modules will indicate how uncertainty in a particular subject propagates within the review to affect its final results. This facilitates the identification of research priorities for better quantifying the links between biodiversity and economics. This is a mechanistic framework, for understanding how changes in biodiversity cause changes in economic costs and benefits, rather than simply for analysing how changes in biodiversity co-vary with economic changes. That is, this framework will help to distinguish causality from simple correlation in the relationship between trends in biodiversity and in economic values. 2.2 Initial considerations and assumptions 2.2.1 Valuing wild nature The term ‘biodiversity’ is often considered to refer to variability amongst organisms (e.g., species diversity, genetic diversity). Indeed, the Convention on Biological Diversity, defines ‘Biological diversity’ as “the variability among living organisms from all sources including, inter alia, terresirial, marine and other aquatic ecosystems and the ecological complexes of which they are part; this includes diversity within species, between species and of ecosystems”. However, the economic importance of wild nature does not rely solely on variability (e.g. species and genetic diversity). Indeed, many of the benefits obtained from nature rely much more on amount (e.g., the abundance of particular species) (Figure 1). Figure 1 Illustration of the importance of variability and of amount for the provisioning of different benefits from nature. a) High variability, low biomass fynbos ecosystem. b) Low variability, high biomass improved pasture. Despite its lower biomass, the fynbos ecosystem has a higher value for the provisioning of benefits that depend on diversity, such as compounds for the pharmaceutical industry. c) High biomass, low variability cod-dominated marine system. d) Low biomass, high variability marine system after overexploitation of cod. Despite its lower variability, the cod-dominated system has a higher value for the provisioning of food, a benefit that is highly sensitive to amount (biomass). Furthermore, the provision of benefits often depends on the condition and extent of ecosystems — incorporating many species and their interactions amongst them and with their environment. Ocean fisheries provision, for example, is affected by the condition of coral reefs and mangroves. It is therefore proposed that the Review on the Economics of Biodiversity Loss should aim to address not simply the effects of the loss in variability, but also the effects of changes in amount, and in the condition and extent of ecosystems. To avoid confusion, throughout we use the expression wild nature to refer to this broader definition of biodiversity. By ‘wild’ we do not mean ‘pristine’ (in which case it would apply to very few parts of the world); we mean ‘non-domesticated’ (i.e. excluding for example livestock and crops). There are many ways in which wild nature contributes to human wellbeing. These have frequently been called ‘ecosystem services’, particularly in the context of the Millennium Ecosystem Assessment (MEA 2005). This term includes both ecological processes (e.g. nutrient cycling, water regulation, pollination) and benefits (e.g. food, water, medicinal products). As discussed in Section 3, benefits are the end products of these ecosystem processes, which directly affect human wellbeing, and which can ultimately be evaluated economically (e.g., clean drinking water). We we focus this Review as much as possible on such benefits, because these can be directly valued. In contrast, the value of ecological processes (e.g., pollination, water regulation) can only be established by valuing their contribution to different benefits (e.g., food crops, drinking water). However, this has not always been possible, and so throughout we refer to benefits or (beneficial) processes as the ways in which wild nature contributes to human wellbeing. This terminology, and how it relates to the framework of the Millennium Ecosystem Assessment, is fully explained in Section 3. 2.2.2 Scope of the Review The aim of this Review is to evaluate the economic consequences of biodiversity (wild nature) loss. It does not address non-economic values of nature, such as those stemming from what many consider the moral right of all species to exist, irrespective of people. Similarly, ecosystem functions (e.g., nutrient cycles, species interactions) are not valued in their own right (or because they are important for other species), but specifically for the direct or indirect benefits they provide to humans. For the purposes of this Review, only changes in wild nature that have economic consequences are considered. Hence, the Review — being an economic analysis — will only be able to cover a fraction of the value of wild nature. There are limits to the meaningful valuation (in monetary terms) of nature and its total complement of services and consequent benefits to humans (Turner 1999; Turner et al. 2003). While most use values can be adequately captured with the economic calculus, controversy remains over the evaluation of so- called non-use values (Pearce & Turner 1990). For some analysts, intrinsic values in nature cannot be ‘captured’ within economic analysis (Pearce & Moran 1994) but much depends on the precise definition of intrinsic value. If the interpretation is one involving the assignment of intrinsic values by a human to nature (or to its components) then techniques such as contingent valuation and choice experiments may be able to shed light on the motivations and preference values involved (known as ‘existence’ value). On the other hand, if the interpretation is one in which no human agent is involved then monetary valuation is replaced by moral imperatives (often linked to non-human species rights and/or interests) (Norton 1992). A further complication arises if one accepts that a certain minimum provision of ‘healthy’ functioning ecosystems is essential to ensure a sustainable flow of services and benefits to avoid threshold effects and system collapse. There is then a ‘glue value’, ‘insurance value’, or ‘infrastructure value’ bound up with this (often unknown) minimum ecosystem provision. Total ecosystem value will therefore always be greater than total economic value (Turner 1991). This document does not reflect our personal conviction, shared by many, that wild nature has an intrinsic value which on its own justifies efforts to conserve it. However, an evaluation of the economic value of wild nature is not incompatible with this conviction. Indeed, if the results of such evaluation are that conservation results in a net economic gain, then that simply adds an economic argument against wild nature loss, alongside the moral argument. If the results are that conservation of wild nature incurs a net economic loss, then that will provide the net size of the bill for conserving wild nature. Note that ‘economic’ does not mean simply monetary revenue. It also includes aspects such as changes in livelihood conditions, economic structure, investment risk and social aspects such as poverty, inequality in access, and benefit sharing. 2.2.3 Valuing marginal change This Review is not about the economic value of wild nature as a whole. In one sense, that calculation is trivial: biodiversity has an infinite value because no human life is possible without it (Toman 1998). This Review is instead about the economic consequences of a marginal loss of wild nature, which are substantially lower than the total stock of nature. Indeed, even within the current biodiversity crisis we are (fortunately) very far from erasing all life and the benefits we derive from it. Even substantial transformations in natural habitats may result in only relatively small changes in the provision of some benefits (e.g. water regulation benefits can be retained in the absence of forest cover through appropriate soil conservation practices; Bruijnzeel 2004) and in some cases changes in wild nature may actually improve the provision of benefits (e.g. it is possible that wild meat production is higher in some secondary habitats than in primary forest; Robinson & Bennett 2004). It is also fundamental to understand that for some benefits the value of nature is to contribute to the provision of a benefit, rather than to provide the entire benefit. For example, a degree of water purification would take place even in the absence of life, as water can be filtered by passing through soil and rock. It would therefore be incorrect to attribute the value of all water purification in a watershed to its natural vegetation. What is relevant for the purposes of this Review is the added (again, marginal) value of natural vegetation to water purification. The concept of marginality is key for making ecosystem service research policy relevant because it is at the margin that policy and economic decisions operate (Turner et al. 1998). Indeed, it would not be useful, for example, to calculate the total value of the global forests as a tool for informing practical forest policy (Bockstael et al. 2000; Fisher et al. in press-b). Having said this, the term ‘marginal’ is usually used in the economics literature to refer to a very small change. Here we refer to changes that while small in relative terms (compared with the full stock of wild nature) may be quite substantial in absolute terms (e.g., thousands of hectares of change in land use). This creates substantial challenges to economic valuation, particularly if the system function is subject to abrupt, non-linear changes in function (e.g. by ‘flipping’ from one equilibrium state to another). Incorporating marginality in ecosystem service evaluation requires therefore a good understanding of the drivers and pressures on the systems under study, as well as of how the system is changing or might change from its current state into a different state under a particular policy action (Fisher et al. in press-b). Here, we use the term ‘marginal’ to refer to changes that a single policy level can foresee. Hence, the valuation of marginal changes is anchored into specific ‘states of the world’ generated from counterfactual scenarios where a specific policy action is either adopted or not (see below, and section 2.3). The overall economic consequences of a marginal loss of wild nature can be evaluated by investigating the economic consequences of adopting a specific set of actions that can prevent that specific level of marginal loss. Such evaluation requires comparing different ‘states of the world’ — one with more, and one with less wild nature. The ‘state’ of a given area (e.g.: a | km plot in Europe; the entire world), is a particular set of biotic and abiotic conditions for that area, comprising aspects such as: land cover, climate, human distribution, human activities, and conservation actions in place. There is a current state of the planet, a different state existed 20 years ago, and there will be a different state in the future. Scenarios can be used to conceptualise the state of the planet in the future (e.g., by 2010 or by 2050) or what it would have been today if decisions in the past had been different. Scenarios are the tools (e.g., models, storylines) used to imagine states of the world. Conversely, states of the world are the outputs of particular scenarios. As discussed in section 2.3.2, measuring the consequences of a marginal change in wild nature requires contrasting very specific states of the world (counterfactuals), in which everything else is equal except for the implementation or not of a set of actions aimed at reducing losses in wild nature. This has the advantage that it requires being explicit about what the conservation goals are, and what the actions are required to achieve those goals. 2.2.4 The need to be spatially explicit A key characteristic of this framework is that it is spatially explicit. We think this is essential, for three sets of reasons: 1. The production, use and flow of benefits from wild nature vary spatially, and so it matters to human wellbeing where conservation actions are implemented. 2. A spatially-explicit framework requires stating the assumptions being made when extrapolating across heterogeneous landscapes using limited data. 3. A spatially explicit quantification of benefits and costs allows makes explicit the mismatch between winners and losers in different scenarios, and is thus essential for designing effective and equitable policy interventions. We develop these points in detail in section 2.3.3. 2.3 The conceptual framework The proposed conceptual framework integrates ecological and economic knowledge to evaluate the net socio-economic consequences of policy actions for halting/reducing wild nature loss (Figure 2). This framework is therefore a practical tool for evaluating the effectiveness of different policy actions. The next section details, step-by-step, how the framework works. SSO] SINQLU PIM Suronpos/suypey 10j suoroe Aorjod Jo soouonbosuoo s1ur0u0d9 pue [eoIGOTOSS oy} SuNENeAdo 1oJ YIOMOUTeIT [enjdoouo0s YW Z IANSIY SSO} auNjeu pyIM vonpalpAjeYy O} uo}jOe JO SaDUanbesuod ASI10d IIWOUGCIS-O1IOS S}EN|eAS eAgigeys ésjeoB uononpe. 2dq9 PX: [ata] ee. jeuojGas yooye Ayianod yoaye jeuopjeu yoaye 1eqol6 yoaye s ysa0op Moy ysa0p Moy 3 Sa0p moy 31 Sa80p Moy ‘Aa we ccc ccoe,. \SI4 pue Moy uo}j0e }yauaq ul SaouasayIp JO anjea s1WOU09a jo sjsoo dew pue Ajyuenb dew pue Ayjjuenb = Se) g uoleurjod doso dew pue Ajjjuenb 'g UOISIAOJd Ja}EM // uoyse dew pue Ajjuenb qm uoneurjjod doJo ul saquasayip JO ANjeA d}WoUs8 dew pue Aj\juenb uoneui)jod dou ul seouaieyip dew pue Ajjuenb @ Salaysly 0} YSU 7] P|JOM dew pue Ajjuenb g@ seveysy aujeur Ajddns Jayem dew pue Ajjuenb ul saquasey!p dew pue Ajjuenb dns sayem u| saouaay/P JO anjen d1lW0u0D8 dew pue Ajjuenb $So| Sunyeu pM aonpaujAjey 0} uoN9e Adijod weeeceereccece? jonoNposd seueus! uonyonpod ertesecosocec, VW uoljeul|jod dois 0} SJ U) SAdUaJayIP ‘ sauaysiy 0} $0 Anjen o1lWOUCD8 yS|1 Ul S2QUBIAYIP dewi pue Ajijuenb dew pue Ajiyuenb dew pue Ajijuenb T UOISsIAold Jayen" Uoljoe t dew pue Ajjuenb JNOUIM salaysl V S@Uaysi O} ¥Si tW P|AOM ul SaouaJeyip dew pue Ajjuenb dew pue Ajquenb SBAYSI} Ul SEOUBJOYIP jO anjen WOUOIS dew pue Ajijuenb sso| anyeu pjim Jo SIaALG Y Saleysi aulseur dew pue Ajijuenb Throughout we refer to a ‘policy action’ but the framework can be used to test packages of policy actions, as long as adequate counterfactual states of the world are created accordingly. The term ‘wild nature loss’ is used throughout to refer to losses in biodiversity (both variety and amount) and ecosystem degradation (section 2.2.1). 2.3.1 Defining appropriate policy actions based on the drivers of loss Biodiversity is being lost and ecosystems are being degraded through a diversity of drivers, including habitat loss and degradation, overexploitation, species invasions, and climate change (Baillie et al. 2004). Drivers of wild nature loss { The starting point of the framework needs to be a good understanding pa of these drivers. This is crucial to desinging and costing effective policy actions for reducing/halting biodiversity loss and ecosystem degradation (losses in wild nature). policy action to halt/reduce wild nature loss oe : : ; 9 For example, overexploitation is the main driver of losses in the weeccccess =’ provision of marine fisheries (Pauly et al. 2005). Appropriate policy actions for reducing/halting the decline in the provision of benefits from marine fisheries address this driver directly by regulating fishing effort. This may include: ereccecocoery Seececeeececec? - Regulation of the temporal distribution of fishing effort by setting fishing seasons that minimise impacts on the fish stocks (e.g., avoiding the reproductive season). - Regulation of the spatial distribution of fishing effort (e.g. creating marine protected areas with no-take zones, banning of bottom dredging practices below 1000 m). - Regulation of overall fishing effort, by setting fishing quotas. - Regulation of fishing targets and impacts, by defining which fishing gear can be used. The first, and crucial, step in the framework is to make explicit what the overall conservation goal is. For example, it may be “to prevent any additional fishing stock from becoming depleted and to ensure the recovery of at least 30% of currently depleted socks by 2030”. This step is crucial because this is where it is defined what specific marginal change in wild nature is being considered in this evaluation (see section 2.2.3). The economic consequences (both costs and benefits) of changes in wild nature (biodiversity loss and ecosystem degradation) are completely determined by this definition. Having defined this goal, the next step is the identification of an appropriate bundle of policy actions, explicitly enough to be costed, that are likely to deliver the goal. For example, these may include actions such as: - Decommissing 80% of the tonnage in fishery X; - Set aside 10% of region Y as reserve networks; - Replace current fishing gear W with new gear Z. Naturally, the definition of what the appropriate actions are needs to be done in close collaboration with experts in the field. 2.3.2 Defining the counterfactual states of the world Evaluating the effectiveness of a particular policy action (or bundle world A: of actions) requires comparing two hypothetical states of the world: no action (business- as-usuay 7 - World A (business-as-usual), where the action is not put in ) place. - World B_ (biodiversity-friendly) where the action is implemented. The states of the world are obtained through appropriate scenarios. The business-as-usual scenario generates predictions for the : : plausible state of the world in the absence of the specific bioaieean intervention being tested (e.g., likely situation of the fishing stocks -friendly) and the fishing industry by 2030 given current predictions in / population growth, in demand for fish, in climate change, etc.). The biodiversity-friendly scenario needs to be identical in everything else to the business-as-usual scenario except but for the specific policy action being tested (e.g. likely situation of the fishing stocks and the fishing industry by 2030 if the policy actions are implemented, given current predictions in population growth, in demand for fish, in climate change, etc.). world B: The scenarios (and the states of the world produced) need to have the right level of information at the right spatial scale. For example, if the action is to implement new protected areas, scenario A needs to state where and roughly how large they are: if the action is to manage fisheries appropriately, scenario A needs to spell out how that would be done, for example by creating no-fishing zones and setting sustainable fisheries quotas. If the scenarios do not have the necessary detail, they will not provide clear answers about how plausible changes in wild nature affect the benefits we derive from it, and they cannot be costed appropriately. While we refer throughout to ‘states of the world’, this framework can be applied to evaluate policy actions at any scale, not just global. For example, it may be applied to evaluate the consequences of implementing a particular protected area or a new dam. The methods for comparing states with and without particular interventions are well-established in the field of Environmental Impact Assessment, being for example a requirement in the EU Directive on Environmental Impact Assessment. Last, while we refer throughout to “two scenarios”, many scenarios can be developed and contrasted, although each contrast needs to be between two appropriately matched counterfactuals. The risks of contrasting inappropriate states It is fundamental that the scenarios are identical in everything else but the specific policy action being tested. Otherwise, the economic results cannot be directly attributed to a 11 difference in the state of wild nature and therefore are not a measure of the economic consequences of biodiversity loss and ecosystem degradation. For example, a comparison between the state of the world today and a predicted state by 2030 is not appropriate for understanding the net economic consequences of changes in wild nature over that period, because many other conditions would be changing at the same time (e.g. population growth, climate change, technology). 2.3.3 Quantifying and mapping how the biophysical provision of benefits is affected by the policy action quantify and map us isheteS : Sy quantify and map quantify and map difierences in world A: a, risk to fisheries A fisheries — without ‘quantify and map action water provision A, Guantify and map This is the focus of Task 3 in this Scoping the Science project (Sections 3 and 4 of this report). The first step is to define what are the benefits or beneficial processes that will be evaluated - i.e. the different ways in which wild nature contributes to human world wellbeing. These need to be defined with uanily and map carefully to avoid double-counting, action water provision B y 4 —— which would compromise the results and ee the credibility of the overall evaluation. Section 3 of this report presents in detail our proposal for a classification of the links between nature and human wellbeing that avoid the double-counting problem and so provides a sound basis for economic valuation. For each particular benefit (e.g. fisheries) or beneficial process (e.g. water regulation; see section 3) one needs to understand how its provision is affected by the policy action (i.e. what are the marginal benefits of the action). In order to do that, we need to be able to quantify the predicted provision for each of the states of the world considered. For example, we need to predict fisheries production under each scenario; these can then be compared to understand how the policy action is likely to affect fisheries. The importance of being spatially-explicit It is crucial that the contrast between benefit provision under different scenarios is done not simply in terms of the overall global value, but by attempting to quantify and map the spatial variation in the provision of the benefit or beneficial process. Indeed, a key characteristic of this framework is that it is spatially-explicit. This is essential, for three sets of reasons: i. The production of benefits or beneficial processes varies spatially, because it is based on spatially-variable underlying ecosystem processes. For example, fisheries production will depend on ocean productivity (section 4.6), while carbon storage (contributing to carbon regulation; section 0) depends on the vegetation cover (Figure 4). The flow of benefits from the point of production (e.g., an upstream forest) to the point of use (e.g., a downstream city) is highly influenced by spatial constraints. The use of these benefits is spatially heterogeneous too, depending critically on the patterns of distribution of end users (e.g., cities, or agricultural areas), which creates substantial spatial variability in the value of benefits even within areas with similar natural production and flow. Finally, alternative states of 12 the world capturing different political interventions need to be spatially explicit, as the results of such interventions depend critically on where they are implemented (e.g., the location of no-take marine areas). Integrating all these components (production, flow, use, and state) clearly therefore requires a spatially explicit framework. ii. Frequently, available information will be insufficient for the elaboration of detailed maps of particular elements of the framework (e.g., value of wetlands for storm protection). This does not render the framework less useful: quite the opposite. Having to describe each element spatially requires stating the assumptions being made when extrapolating from limited data. If a single data point is all that is available, then the corresponding world map may have a single colour. Oversimplifications become obvious, and so are more likely to receive attention in future work. Further information will help to refine the spatial representations (e.g., if other data indicates that value of wetlands for storm protection is likely to vary with Gross Domestic Product/km’). A spatially-explicit framework can always be aggregated into coarser spatial units, for example to provide values per biome, or global values. The reverse, however, is not possible. iii. Services often flow from their point of production to users elsewhere (Figure 5) — so the benefits of their conservation may accrue to different actors than the benefits associated with their loss (Figure 6). For any given action there will likely be winners (e.g., areas where fisheries production increases) and losers (areas where production decreases). Information on these would be lost from the global aggregated value, but is fundamental to an appropriate economic valuation (as in most cases benefit value is context-dependent). A spatially-explicit approach is therefore crucial to fully evaluate the broader social consequences of each action in terms of impacts on livelihoods, development goals, and equity, and for designing effective and equitable policy interventions. Quantifying and mapping the production of benefits/processes Obtaining global, or at least large-scale, maps of the production of a particular benefit (or beneficial process) requires a good understanding the factors that drive production. These factors will include a mix of abiotic (e.g., climate, soil type, topography) and biotic (e.g. ecosystem type, species diversity) variables, as well as factors that are determined by human actions (e.g. extent of habitat). If these relationships are well understood, it should be possible to generate a mathematical model that can reasonably predict benefit production given information on the underlying factors. Such a model corresponds to what in economic terms is called a production function. At this stage, however, we are interested in modelling the biophysical production of benefits (or beneficial processes) rather than their economic value. For example, timber production from natural forests per area, per year, may be given as the forested area multiplied by a function (f) of factors such as primary productivity, forest type, topography, etc. (Figure 3). 13 ® | | | timber production = forest area x primary productivity, forest type, tree diversity, topography, ....) | | & | | Figure 3 Hypothetical production function for timber form natural forests. Green arrows correspond to biodiversity/ecosystem inputs; red arrows correspond to human inputs. Note that for many of these benefits their production depends not only on wild nature, but also on abiotic (e.g. topography) and anthropogenic inputs (e.g. forest area). The main aim of this Scoping the Science project is to understand the marginal (or added) value of wild nature for the production of these benefits. Given such a production function and appropriate biophysical data (maps of primary productivity, of forest types, of tree richness, topography) it would be possible to create a global surface of potential timber production for each state of the world. From one state to the other, some key variables would change (e.g., forest conversion), therefore resulting in spatial differences in timber production. A contrast between these maps would then be used to quantify, and subsequently value, the spatial variation in differences in timber production between states of the world. The production map should be expressed in the appropriate physical units for measuring the wile . . . . io) provisioning of the service in question (e.g., tons of carbon/km? for carbon storage, m* of 2) opis. . oe 2 . . water/km*/y for water provisioning, numbers of visitors/km*/y for recreational services). In practice, a detailed model of a production function will seldom exist at a global scale. However, more simplified models do exist in some cases, or they can be produced from available data. For example, a global model of wild meat production (section 4.8) has not been created yet, but a reasonable number of individual studies have calculated productivity for different areas around the world, and we believe that a first-cut global model could be produced in time for Phase 2 (section 4.8). The key question we addressed in each of the thematic reviews (section 4) was how far is scientific knowledge from being able to create such models. In a few cases, maps of benefit/process production have already been developed (e.g. Figure 4). 14 Figure 4 A preliminary map of the global distribution of current ecosystem productivity for the service ‘carbon storage’ (Naidoo et al. in press). Mapping (potential) sustainable production, rather than current use For some benefits (those that involve harvesting, such as fisheries, timber production, wild meat production) there is a potential for overexploitation leading to depletion. In this case, the relevant production map needs to correspond to a map of sustainable productivity, rather than current benefit flow. For example, when considering food production from fisheries, the relevant map is one of sustainable fisheries production, rather than one of current catches. Indeed, the latter is likely to reflect excessive fishing effort in some areas, which over the long-run will lead to a reduction in benefits. On the other hand, a map of current benefit flow would ignore areas where there is no present use but which may be valuable nonentheless. For example, areas too remote to contribute for current timber production may be considered to hold value for future timber (option value, Section 3.7). Quantifying and mapping risk Biodiversity loss and ecosystem degradation affect not only the flows of services and benefits, but also the resilience of systems. It is therefore crucial to attempt to quantify the extent to which a particular action (or the lack of its implementation) affects the likelihood that each service or benefit will be compromised. Knowledge on resilience and risk is even scarcer than on benefit/process production, but whenever available is should be incorporated in the economic valuation. In the thematic reviews (section 4), we collated information on the possibility of non-linearities in the production of each benefit/process, and particularly the possibility of threshold effects in which a small change in the state of wild nature could result in a disproportionate change in benefit/process provisioning. For example, there is now some information on the conditions that trigger collapse in marine fisheries (section 4.6), and so it may be possible to obtain at least some indication of the extent to which fisheries are at risk of collapse under each scenario. Again, whenever possible, this information should be quantified in a spatially-explicit way, even if the mapping units are likely to be quite coarse. For example, it will clearly not be possible to map risk of fisheries collapse at a fine scale, but expert opinion could probably provide at least a qualitative assessment of the risk of collapse for each of the FAO major fishing areas. Key here is considering the sensitivity of benefit/process productivity to likely changes in biodiversity. It may not be vital (at this stage) to know, for instance, that reducing forest cover from 2% to 1% of the land surface will probably trigger a collapse in timber productivity if the relevant difference in states of the world (under policy action vs. inaction) is 25% vs. 15% forest cover. 2.3.4 Quantifying and mapping the economic value of changes in benefits derived from the policy action The information on how the biophysical production of each _ benefit/process changes under each scenario becomes the basis of the economic valuation, using appropriate tools. quantify and map economic value of differences in fisheries quantify and map differences in fisheries Guantify and map differences in risk to fisheries production quantify and map economic value of differences in risk to isheries productio It is beyond the remit of this project to discuss in detail how such valuation can be made, but we present here some considerations that will affect the way the valuation is done. quantify and map differences in water supply quantify and map economic value of differences in water suppl. Different types of values quantify and map economic value of differences in crop pollination quantify and map differences in crop pollination Different benefits (or beneficial processes) correspond to different types of economic values, which affects the methods used to os 33 = es <= 2 2 s a a c Sec ‘a oO = s S iH ob 3 > E e § = o§ es a a 2 o 2 ca co ° = D tc o- =e a ie os 3 ny og See oie 2 td € o— oe 2 =e oe ap £S So §3 = 23 = 1S. ne oo =f = o> = 8 Ce i= c ao Sse = A {Ss as ag = ca so ose ¢ > mas so 26 86 2 s Ss ae os 5 52 <2 ~£ ~¢ ss 23 s£) 53) 5 | s | g¢ | Sa) 22 Se) S | eo) 8 | ER Es | 2 | se | Se |) 33) si) alee a — — ° = = ° of = = = , = oo o8 2 a GE Le as Of = a a 2a | ea q es 2S 385 35 oe 56 Crops Livestock Marine fisheries Inland fisheries Wild animal products Drinking and industry water Hydroelectric energy Wild plant fibers Wild medicinal plants Nature-related outdoor activit Avoid. injury, property loss One-off use benefits Non-use benefits Unknown benefits Figure 10 Review themes in this project (1 to 17), and how they relate to beneficial ecosystem processes (columns) and benefits (rows). See text details on each number. Grey bars on the left indicate main types of benefit. Each review theme is represented by a block with a blue border. Colours indicate an estimate of the importance of the link between each ecosystem process and each type of benefit: red — very important; orange — important; yellow — somewhat important; white — relatively unimportant. 37 For example, we considered global climate regulation (including regulation of ocean acidity) as a single (‘vertical’) theme, rather than considering it within each of the relevant benefits (e.g. a part of crop production, of avoidance of injury and property loss, of the production of marine fisheries; Figure 10). We decided to do so because the Stern Review (Stern 2007) has, effectively, already valued the consequences of avoiding climate change by considering its effects on each of a diversity of benefits, including food crops, injury and property damage, and marine fisheries. Stern did not include all the possible links between climate change and human wellbeing (e.g., how climate change affects recreation, tourism, research and education, spiritual and cultural wellbeing), but we recommend that it would be a better investment of resources in Phase 2 to build from the Stern Review rather than to repeat their analyses (and invest those resources in any of the other themes, which are much less advanced in terms of valuation). In order to avoid double-counting, the effects of climate change on benefits have been excluded from all the other themes (Figure 10). 3.6 Thematic reviews Overall, we considered 17 thematic reviews (Figure 10). Some of these have the same titles as the benefit categories listed before, but often the review had a narrower focus/definition, as explained: e Wild crop pollination (1): how wild nature contributes to crop yields through wild (unmanaged) pollination (section 4.1). e Biological control in crops (2): how wild nature contributes to crop yields through the control of crop pests and diseases (section 4.2). e Genetic diversity of crops and livestock (3): how wild nature contributes to agriculture and livestock production by maintaining a diversity of populations of wild relatives and varieties (section 4.3). e Soil quality for crop production (5): how wild nature contributes to crop yields by contributing to soil formation and to preventing erosion regulation (section 4.4). e Livestock (5): we focused specifically on the way wild nature contributes to livestock production through the provision of natural grazing and browsing areas (rangelands) (section 0). e Marine fisheries (6): we focussed specifically on capture fisheries (mainly for food), with some considerations on recreational fisheries (section 4.6). e Inland fisheries (7): we focussed on capture fisheries and aquaculture (mainly for food), with some considerations on recreational fisheries (section Error! Reference source not found.). e Wild animal products (8): we focused on wild meat, with some considerations on recreational hunting (section 4.8). e Fresh water provision and regulation (9): how wild nature affects freshwater quantity, timing and quantity (section 4.9). 38 Wild plant fibres (10): we focussed on timber production (section 4.10). Wild medicinal species (11): how wild nature contributes to human health through the provision of a diversity of harvestable medicinal plants (section 4.11). Nature-related outdoor activities (12): how wild nature contributes to benefits such as human health, psychological wellbeing and education through outdoor activities such as nature tourism, hiking, cycling, diving, camping, species- watching, use of urban parks, and gardening (section Error! Reference source not found.). Natural hazard regulation (13): how wild nature contributes to avoid human injury and property loss by avoiding or mitigating the effects of natural hazards, including coastal storms, hurricanes, floods, avalanches, mudslides (section 4.13). One-off biodiversity use values (14): how wild nature contributes to health, psychological wellbeing, physical comfort, and knowledge by maintaining the biological diversity from where ideas, chemical compounds, and images can be sourced. We focused on pharmaceutical compounds (section 4.14). Non-use values (15): non-material benefits, in which appreciation of nature (species and landscapes) results in improved psychological wellbeing, knowledge and social relations without a direct use of biodiversity. These may translate into cultural diversity and heritage, spiritual and religious values (inc. sacred species and groves), knowledge systems, educational values, aesthetic values, social relations, and sense of place (MEA 2005). The benefits they deliver are mainly in terms of psychological wellbeing (e.g. sense of wonder) (section 4.15). Global climate regulation (16): how wild nature contributes to human wellbeing by contributing to climate regulation (concentration of greenhouse gases in the atmosphere). We focused on terrestrial systems (section 0). Unknown benefits or processes (17): wild nature may potentially contribute to our future welfare in ways currently not realised, including both through ecosystem processes not currently known or valued (e.g., carbon storage and sequestration has only recently been identified as a ecosystem valuable process) and through benefits currently not predicted (e.g., algae are just starting to be considered as a promising biofuel). While these benefits are difficult or even impossible to measure, they may be very substantial (section 4.16). In addition to these main review themes, we considered three cross-cutting themes: Resilience: review the evidence for how wild nature contributes to the resilience in the provisioning of each of the benefits and services described in points | to 17 (this was covered under each main theme). Scenarios: brief overview of main tools available for generating scenarios that could be useful in Phase 2 (section 4.17). Under each main theme there is also a brief description of the main type of information that scenarios would need to produce. 39 e Prioritisation: a systematic assessment of the relative priority of the recommended analysis under the main themes for Phase 2 of the review (and beyond) based on a combination of predicted feasibility (given the state of ecological knowledge) and importance (how they are likely to affect the results of the valuation) (section 4.18.1). 3.7 Relationship between the proposed classification and the Millennium Ecosystem Assessment framework The MEA followed a classification of ecosystem services into supporting, provisioning, regulating and cultural (Figure 7). These types of services map into our proposed classification as illustrated in Figure 11. Essentially, the MEA’s supporting services correspond mainly to our ‘core’ ecosystem processes, although they also include some ‘beneficial’ services. The regulating services essentially correspond to our ‘beneficial’ processes, and provisioning and cultural services tend to correspond to our benefits. The match is not perfect: the MEA provisioning service “genetic resources’ (underpinned by our processes ‘genetic diversification’ and ‘species diversification’) is spread across a diversity of benefits such as crop and livestock genetic diversity contributing to food production, diversity of wild pollinators contributing to crop yields, and diversity of wild species contributing to tourism. The MEA regulating service ‘natural hazard regulation’ is partially covered by our process “fermation of physical barriers’, but also by water, climate and erosion regulation. The benefits provided by these services include the avoidance of personal injury, of damage to property and infrastructure, of damage to crops, and of damage to tourism. 40 BENEFITS BENEFICIAL ECOSYSTEM PROCESSES CORE ECOSYSTEM PROCESSES Biological cont Other ecological interactions e Formation of species habitat o Working animals o Hydroelectric energy e Weathering/ erosion ° e Pri i e Ecological interactions ell e Evolutionary Physical health processes clit julati e Currently unknown beneficial processes Psychological wellbeing Correspondence with the MA classification ee) Supporting services eee Provisioning services {HNN Provisioning services (partial match) Rest Regulating services A Regulating services (partial match) . ee Cultural services K Currently unknown benefits Figure 11 Correspondence between the classification followed in this review (into core ecosystem processes, beneficial ecosystem processes, and benefits) and the classification followed by the Millennium Ecosystem Assessment (into supporting services, provisioning services, regulating services, and cultural services). 4] 3.8 Relationship between the proposed classification and the Total Economic Value framework While the MEA is still relatively recent, and few economic papers have been published on it, a longer and well-established approach for classifying the value of wild nature is the Total Economic Value (TEV) framework (Pearce & Turner 1990, Pearce & Moran 1994; Figure 12). Here we clarify the relationships between the review tasks and the TEV framework (Figure 13). Throughout, we use the definitions in Defra (2007) for each type of value. | Total Economic Value Non-use value RS SENTEN ES epson i For others Existence | use value Direct use | Indirect use Altruism Bequest | J ges ee Figure 12 Total Economic Value framework (adapted from Defra 2007). Direct consumptive use values Direct consumptive use values are those obtained when individuals make actual or planned use of an ecosystem by extracting resources from it. Amongst the benefits/processes we are reviewing, this corresponds to those where there is direct harvesting from nature: livestock grazing in natural areas (theme 5), marine and inland fisheries (including recreational fishing; 6 and 7), harvesting of wild animal products (including recreational hunting; 8), harvesting of wild plant fibers (10), and harvesting of medicinal plants (11). Direct non-consumptive use values Direct non-consumptive use values are defined as being obtained when individuals make actual or planned use of an ecosystem without extracting any element from it. Nature tourism and other forms of outdoors recreation related to nature are traditionally considered within this category (e.g. Rockel & Kealy 1991), and we followed that approach (theme 12), although it should be acknowledged that these activities may have substantial impacts on the places visited, even if not harvesting resources from them. c s = s <= = 5 a = = 5 z Blomass production: primary Biomass production: secondary Biological control Other ecological Waste agsimilation Erosion regulation physical barriers pleasant scenery Water regulation (timi Interactions species habitat Pollination Formation of diversification Genetic diversification Soll formation Formation of Formation of Air purification (quantity) regulation Unknown processes Spec! "|| Water provisioning ~— || Global climate = a= —— Livestock Marine fisheries Inland fisheries Wild animal products Orinking and industry water { Hydroelectric energy 17 Wild plant fibres Wild medicinal plants Nature-related outdoor activit Avoid. injury, property loss Non-use benefits Unknown benefits 17 | Total Economic } Non-use Option Figure 13 How each main review theme relates to the Total Economic Value framework. Indirect use values Indirect use values refer to situations when individuals benefit from ecosystem processes underlying the provision of direct benefits. For example, crop pollination (theme 1) is not used as a benefit per se, but as a process towards the increased provision of crop yields. We also included in this category: biological control in crops (2), maintenance of soil quality for crop production (4), fresh water provision and regulation (9), global climate regulation (9), and the regulation of natural hazards (13). Option values Option values are the values that people put on the option to use a resource in the future, even if they are not current users. These future uses may be either direct or indirect. Option values can be thought of as a form of insurance, but the economic literature is divided over 43 whether such values are a true component of TEV (Freeman 1993). Strickly speaking, we should treat them as a separate concept related to a precautionary policy stance (theme 17) Wholly within option values is genetic diversity of crops and livestock (theme 3), which can be considered a form of insurance in the case of future need to add genetic diversity to existing crop varieties or livestock breeds, for example in case of emergence of a new pest or disease. Also within option values are benefits from one-off use (14), the set of the ideas, chemical compounds, and images initially obtained from nature but subsequently propagated outside of it. Indeed, by definition the only way in which wild nature is valuable for these is through predictions of their future value. Data on the economic value of previously extracted ideas, compounds and images can be used to obtain estimates of likely future value, but not for measuring directly the value of existing wild nature (as those past ideas, compounds and images are now fully independent from nature). All tasks covering use values (direct and indirect) listed previously can potentially have a component of option value. For example, not all known marine fish stocks need to be exploited today (even sustainably). One possibility is that some stocks may be exploited immediately and some left untouched as options for the future. These two types of values are likely to be monetised differently in the subsequent economic valuation. As an example of option values for indirect use, natural water purification processes have no use value if nobody is using the water, for example in a region with no inhabitants. Nonetheless, we may want to attach value to those processes (and the ecosystems generating them) as providing options for future use. Option values also apply to non-consumptive uses: we may want to set aside areas as valuable for future tourism, even if currently they are not used as such (e.g. deep sea systems may become attractive for tourism given future technological developments). All considerations of resilience discussed within each task refer to option value, as they relate to the concern of ensuring future use. For example, the diversity of marine functional groups provides resilience to fisheries production (see section 4.6), and so there is value in the conservation of such variety as a form of insurance against future changes. Non-use value Non-use value is derived simply from the knowledge that the natural environment is maintained. There are three main components: - Bequest value: where individuals attach value from the fact that a given natural resource is passed on to future generations; - Altruistic value: where individuals attach values to the availability of a given natural resource to others in the current generation; - Existence value: where individuals attach value simply from the existence of a resource, even though the individual has no actual or planned use of it. Theme 15 (non-use values) relates principally to existential non-use values. Bequest values — which directly address issues of intergenerational equity — are implicit or explicit in other aspects of the Review on the Economics of Biodiversity Loss. In particular, 44 we recommend that the comparison between states of the world for a given resource (e.g. fisheries production, wild meat) is based on the capacity of ecosystems for the sustainable production of the resource, rather than based on current flows that may not be sustainable (see section 2.3.4). In valuation terms, bequest values and the associated concerns about intergenerational equity will be explicitly incorporated in the decision of which discount rate to apply when valuing future resources, as was also the case with the Stern Review on the Economics of Climate Change (Stern 2007). They can also be addressed in the estimation of values (based on contingent valuation methods). Altruistic values, on the other hand, include (amongst others) concerns for intragenerational equity. The emphasis placed on the spatial distribution of costs and benefits (see section 2.2.4) is aimed at addressing such concerns, allowing for the redistribution of benefits for example through the establishment of payments for ecosystem services. Some of these are not altruistic, though, as such redistribution may directly benefit the ‘donors’ by increasing the likelihood of long-term provision of the resource or by reducing the deleterious social effects of inequity (for example in terms of security). Purely altruistic values nonetheless exist and can be integrated in the valuation component of the Review. But the only way to assign monetary values to non-use motivations is via so-called stated preference survey/choice-based methods. It is still an open research question as to how adequately values such as existence value can be meaningfully expressed in monetary terms (Heal 2000; Turner et al. 2003; Barbier 2007). To the extent that non-use values are not captured via economic analysis the TEV and the Total Ecosystem Value (TSV) will diverge with TEV < TSV. 3.9 Participants Authors Andrew Balmford, Ana Rodrigues (University of Cambridge) Rhys Green (University of Cambridge and Royal Society for the Protection of Birds) Kerry Turner, Brendan Fisher (Centre for Social and Economic Research on the Global Environment, University of East Anglia, UK) Robin Naidoo (Conservation Science Program, WWE-US, USA). Acknowledgements We are grateful to the following for providing opportunities for stimulating discussions in the development of this section: partners in the Scoping the Science project: Patrick ten Brink, Marianne Kettunen (IEEP), Leon Braat (Alterra); Aude Neuville and Patrick Murphy (DGENV); members of the Working Group on the Review on the Economics of Biodiversity Loss. 45 4 THEMATIC REVIEWS Given time limitations, we were unable to fully develop all the themes listed in the previous section. As a result, themes have been analysed at different levels of depth, from fully developed thematic reviews that received expert support, to not developed at all. The level of each review is indicated at the start of their respective section. We tried to answer each of the following questions in each thematic review. However, for the least detailed reviews that was not possible. At very least, we tried to obtain an answer to the questio: Can we quantify and map the global provision of this benefit/process and how it might change? We aimed to evaluate how far scientific knowledge is from being able to produce a spatially-explicit quantification of the production of the benefit (e.g. fisheries production) or of the function of the beneficial process (e.g., added value of wild pollination for crop yields), that can act as a basis for the economic valuation contrasting two states of the world. The information gathered was then used to guide the prioritisation of recommendations for Phase 2 (section 4.18). The questions are: e Why is this benefit/process important for human wellbeing? A brief review of the information supporting the relevance of each benefit/process for human wellbeing. Any monetization presented here is to support the importance of the benefit to human welfare, rather than to quantify the importance of wild nature for the provision of the benefit. For example, the information that “floods affected 3.5 million people in Cambodia (with associated costs of US$145 million) in 2000” should not be interpreted to mean that conserving wild nature could have prevented those floods and therefore avoided those costs. e What are the overall trends in the provision of this benefit/process? Review of state of knowledge on the trends of the provision of the benefit/function. e How is the provision of this benefit/process affected by changes in wild nature? Here we describe the links between wild nature and the production of the process/benefit. This section is tailored to each review task, discussing the most relevant aspects of wild nature in each case. These may include ‘diversity’, such as species richness, and ‘amount’, such as biomass, as well as ecosystem condition. We specifically looked for information on which each of these components is particularly important for resilience in the provision of the benefit/process. And we specifically tried to understand what the relationship is between habitat area and the provision of the benefit/function, which is important for estimating the likely consequences of habitat loss. e What are the main threats to the provision of this benefit/process? Information on the drivers of loss is key to the development of scenarios that are fit-for-purpose (section 2.3.1). Indeed, these will only be accounted for in the Review on the Economics of Biodiversity Loss if they are different in business-as-usual and biodiversity-friendly states of the world. The information on these threats is also fundamental to calculating the costs associated it conservation, as costs will vary depending on the type and intensity of threat (see separate report on “Review of the 46 Costs of Conservation and Priorities for Action”, by Bruner, Naidoo and Balmford, also an output of this Scoping the Science project). Are abrupt changes likely in the provision of this benefit/process? Here we were aiming to understand if there is evidence that the provision of this benefit/ process may be subject to thresholds/tipping points in the foreseeable future (in the questionnaire to experts we used 2025 as the foreseeable future data). By thresholds/tipping points we refer to a situation when a small (anthropogenic) change in nature may have disproportionate effects on the provision of the benefit/process. We drew from the information in the previous sections and on the opinion of experts to make a prediction of the likelihood of thresholds or tipping points. The result is a qualitative (and inevitably subjective) assessment of relative risk of abrupt changes. Can we quantify and map the global provision of this benefit/process and how it might change? Here we evaluate how far scientific knowledge is from being able to produce a spatially-explicit quantification of the production of the benefit (e.g. fisheries production) or of the function of the beneficial process (e.g., added value of wild pollination for crop yields), that can act as a basis for the economic valuation contrasting two states of the world. At one extreme, a good model already exists (a well-developed production function) and maps can easily be generated from available data. At the other extreme, no good model and/or good data exist that could form the basis of a global map. In most cases the state of knowledge was somewhere in between, with at least a first cut being possible within one year by building from existing studies. We also tried to obtain information on: the main gaps in data/knowledge; what would be needed from scenarios for the model proposed; who could potentially do the analysis; and what the required effort would be (in researchers-months). Insights for economic valuation The purpose of this section is to maximise the value of our work as a basis for the economic valuation, helping readers to interpret our recommendations by clarifying what our assumptions were on how the ecological modelling and the valuation fit. Some key resources Here we list key resources that we came across in our review and which are likely to prove useful in Phase 2. Participants The list of authors, contributors, reviewers and acknowledgements (see Annex 3). 47 4.1. Wild crop pollination This section is a fully developed review, including contributions by experts in the field, who subsequently reviewed the text. 4.1.1 Why is this process important for human wellbeing? Many plant species benefit for their reproduction on animals that transport pollen between flowers. This biotic pollination is typically done by insects (particularly bees; also flies, beetles, moths, butterflies and wasps), but in some species is performed by vertebrates (particularly birds and bats). Pollination is therefore a key ecological process service, upon which both natural and agricultural systems depend (Nabhan & Buchmann 1997). While the extent to which staple food crops depend on pollinator services has been questioned by some (Ghazoul 2005), Klein and colleagues (2007) found that fruit or seed numbers or quality of 87 out of 115 leading global crops (representing up to 35% of the global food supply) were increased upon animal pollination. In many agricultural systems, pollination is actively managed through the establishment of populations of domesticated pollinators, particularly the honeybee Apis mellifera. However, the importance of wild (i.e. unmanaged) pollinators for agricultural production is being increasingly recognised (e.g. Westerkamp & Gottsberger 2000; Kremen et al. 2007; Kremen & Chaplin-Kramer 2007). For a diversity of crops, it has been found that wild pollination increases the size and quality of harvests (Klein et al. 2007). Wild pollinators may also interact synergistically with managed bees to increase crop yields (Greenleaf & Kremen 2006). Furthermore, a diverse assemblage of native pollinators provides insurance against year-to-year population variability or loss of specific pollinator species (Kremen et al. 2002; Rickets 2004; Tscharntke et al. 2005), and might better serve flowers because of pollinator-specific spatial preferences to a flowering plant or crop field (Klein et al. in press). Given current declines in populations of managed honeybees (Colony Collapse Disorder [Johnson 2007], and abandonment of beekeeping in regions affected by ‘Africanization’ of honeybees [Brosi et al. 2007]), the importance of wild pollination is likely to increase. Estimating economic value is difficult and controversial, but the global value of wild and domestic pollination has been estimated at $120 billion per year (Costanza et al. 1997), while Losey & Vaughan (2006) estimated that wild pollinators alone are responsible for about $3 billion of fruits and vegetables produced in the United States. 4.1.2 What are the overall trends in the provision of wild crop pollination? The Millennium Assessment indicated a low to medium certainty that pollination ecosystem services are declining (Duraiappah et al. 2005). Direct evidence for the decline in pollination services (i.e., evidence that global crop yields are being affected by a reduction in wild pollination) does not seem to exist, and would be difficult to obtain given that many other aspects of agricultural practices are changing simultaneously. Some indirect evidence exists based on reported declines in the abundance or area of occupancy of some wild pollinators, for example in North America (NRC 2007), and in Europe (Biesmeijer et al. 2006; Goulson et al. 2008). On the other hand, some species have extended their ranges (Ghazoul 2005). A continent-level assessment of pollinator declines 48 is still lacking (Diaz et al. 2005). Most of the evidence for a decline in wild pollinator services is inferred from changes in land use known to affect pollinator communities, particularly declines in the extent and condition of available natural and semi-natural habitats, and the effects of agricultural intensification (e.g. Kremen et al. 2002). On the other hand, an increase in organic agriculture (Willer et al. 2008) may be increasing provisioning of pollination services (Holzschuh et al. 2008) in some parts of the world. 4.1.3 How is the provision of wild crop pollination affected by changes in wild nature? Many crops are self-compatible to different degrees or wind pollinated, and therefore receive only small benefits from wild pollination services (Klein et al. 2007). For crops that rely heavily on biotic pollination, there are several ways in which changes in biodiversity can affect yields. Wild pollinators often depend on natural or semi-natural habitats for the provisioning of nesting (e.g. tree cavities, suitable soil substrates) and floral resources that cannot be found within crop fields (Kremen et al. 2004). Consequently, the available area of natural habitat has a significant influence on pollinator species richness (Steffan-Dewenter 2003), abundance (Heard et al. 2007; Morandin et al. 2007), and pollinator community composition (Steffan-Dewenter et al. 2002; Brosi et al. 2007). Accordingiy, habitat area in the neighbourhood of crop fields has been found to be strongly related to a direct measure of the pollination service measured here in terms of pollen deposition provided by bees (Kremen et al. 2004; Figure 1a). Besides area, the quality of the habitat, both in natural systems and in croplands, seems to be important for pollinator services, particularly the extent to which they provide nesting and floral resources (Klein et al. 2003a; Goulson et al. 2005; Potts et al. 2005). Not all pollinators are dependent on ‘natural’ habitats (Ricketts et al. 2004) and some are able to use resources within agricultural fields themselves (reviewed in Kremen et al. 2007) and therefore can even profit from agricultural management (Klein et al. 2002, Westphal et al. 2003). There is clear evidence that wild pollination is strongly related to proximity to natural or semi-natural habitats. A recent quantitative review of 23 studies (Ricketts et al. 2008) found an exponential decay in pollinator richness and native pollinator visitation rate with distance to natural or semi-natural habitats. A decline in yield was less clear, possibly because few studies measured it directly. Visitation rate declined more steeply (dropping to half at 0.6 km) than richness (1.5 km). Despite the general exponential relationship found by Ricketts et al. (2008), it may be that slightly different relationships apply to different crops (e.g. linear for coffee, Klein et al. 2003a,b; log-linear for watermelon, Kremen et al. 2004) and to different types of pollinators (e.g. of different sizes; Klein et al. in press). Density of wild pollinators at the crop site is considered a good proxy of visitation rates (Kremen et al. 2007; Ricketts et al. 2008) and therefore key for crop production (Vazquez et al. 2005). This is confirmed by a few experimental case-studies (Roubik 2002), although results are potentially confounded by a positive relationship between numbers of individual visitors and diversity of pollinator species (Kremen & Chaplin-Kramer 2007). Interestingly, Klein et al. (2003b) found that numbers of individual visiting bees did not explain fruit set but that diversity of visitor species did. Three potential mechanisms may lead to a relationship between pollinator diversity and pollinaiton services, ‘Sampling effect’ a 49 greater chance of having a pollinator species that perfectly fits the flower morphology leading to more reliable pollination; (2) ‘Niche complementarity’ all receptive flowers over an extended blooming period receive optimal pollination service; (3) ‘Functional facilitation’ synergistic interspecific interactions with domestic honey bees (Greenleaf & Kremen 2006, Klein et al. in press). The identity of the pollinator species matters, as different crops benefit from pollination by different species (Klein et al. 2007). Some crops have specific pollinator requirements, either certain guilds or certain flowering times, and so abundance of pollinator individuals may not translate into high pollination services if they belong to the wrong guilds (so they visit flowers but are inefficient at actually pollinating them), if their abundance is out of synchrony with crop mass flowering (e.g. almonds in California typically flower before many wild bees have built up significant populations) (Ricketts et al. 2006; Kremen et al. in press; Klein & Kremen unpublished), or if honey bees are attracted to another flowering crop adjacent to the target crop field but other pollinating species still prefer to forage at the target crop e.g. for alfalfa pollination.Accordingly, crops with a narrow range of specialised pollinators are more likely to experience pollinator shortage when grown in highly modified landscapes (Klein et al. 2007). Overall, functional diversity (diversity of functional traits) may be more important to crop yield than either pure abundance or species richness (Klein et al. in press). Consequently, the order in which species are lost is likely to affect pollination services differentially (Larsen et al. 2005). Unfortunately, large-bodied bee and beetle species seem to be both most extinction-prone and most functionally efficient, their loss contributing to rapid functional loss (Larsen et al. 2005). A current NCEAS analysis into body size, nesting behaviour, trophic specialisation and sociality will shed more light on this issue. Pollinator diversity has been found to improve resilience of crop production, by buffering pollination against asynchronous fluctuations of bee abundances between years, including temporal variation in the relative abundance within native species (Kremen et al. 2002) and a sharp decline in domestic honeybees (Ricketts 2004, Winfree et al. 2007b). Given that different species are differentially effective as pollinators (both within and among crops), managing for bee diversity could meet the pollination requirements of a greater number of crops, provide insurance in the event of shortages of any specific pollinator (managed or unmanaged), and provide options for new or alternative crops (Kremen et al. 2002). Furthermore, different taxa are likely to respond differently to landscape isolation and habitat characteristics (e.g. Steffan-Dewenter et al. 2002; Klein et al. 2003a,b), so a diversity of pollinator taxa may help to reduce the effects of land use change on pollination services (Ricketts et al. 2008). The current precipitous declines in managed honeybees in the United States due to Colony Collapse Disorder are making clear the dangers of relying on a single pollinator species (Johnson 2007; NRC 2007). Relationship between habitat area and pollination Most services from wild pollinators take place near the interface between natural/semi- natural habitats and crop fields (Ricketts et al. 2008), and so landscape patterns have a strong influence on pollination services. There is strong evidence that both richness of pollinator species and rates of visitation by native pollinators at crop sites decline quickly with distance to natural/semi-natural habitat (Ricketts et al. 2008; Figure 14a) and so crop yields are also expected to decline with distance. 50 » — richens/native visitation 4000 ‘distance 6 nearest vabitat patch ) fraction of habitat within 2.4 km of farm site ‘ pollen deposition on farm site benefits from pollination © on farm site size of habitat patch adjacent to farm site benefits from pollination & across landscape fraction of habitat within the landscape 1 0 51 Figure 14. Four aspects of the spatial relationship between habitat area and provisioning of pollination services. a) Both species richness (blue) and the rates of visitation by native pollinators (red) have been found to decline with distance to nearest habitat patch (Ricketts et al. 2008), and so presumably the benefits from pollination (in terms of crop yields) should also show a strong relationship with distance. b) Given a particular farm site (cross), pollination increases with the fraction of habitat (green) within 2.4 km of the site’s neighbourhood (after Kremen et al. 2004). c) Our prediction for the relationship between benefits obtained from pollination at the farm and the size of the habitat patch (green) adjacent to a particular farm site (purple): very small habitat patches (smaller than a minimum viable area for pollinators) may not have any effect; afterwards, pollination is predicted to increase rapidly with area, but then to stabilise as further increases in habitat happen correspond only to “core” area (dashed). d) Predicted relationship between the fraction of overall habitat area (green) within a landscape (square) and overall pollination services for the crop (purple; Morandin & Winston 2006): pollination values are zero if there is zero habitat for pollinators; pollination then increases as habitat area increases, but declines again as habitat expansion results in a decline in crop land at the interface with the habitat; for 100% natural habitat, pollination for crops has zero value. For a given overall area of natural/semi-natural habitat, pollination benefits may be higher (dashed line) or lower (solid line) depending on the — spatial arrangement of the habitat patches. The fraction of natural/semi natural habitat area in the neighbourhood of crop fields has been found to be strongly related to a direct measure of the pollination service, pollen deposition (Kremen et al. 2004; Figure 14b). Concrete studies on species—area relationships for pollinators (particularly in tropical forest fragments; Priess et al. 2007) are lacking, but patch area of natural habitat is likely to influence species richness and abundance of pollinator communities, as larger habitat fragments in many cases show higher species richness and density (Fahrig 2003). We therefore predict that, after a minimum threshold of habitat patch size, pollination at a given crop site increases with the size of the adjacent habitat patch, both because the patch itself becomes larger (and potentially has higher pollinator diversity and abundance), and because a larger fraction of the site’s neighbourhood becomes natural/semi-natural habitat (Figure 14c). This increase, however, is only expected up to a certain point, followed by stabilisation, as additional area increases take place too far from crop fields to produce pollination services (Figure 14c; however, Brosi et al. [2007] found non-significant relationships between bee abundance or diversity and size of neighbouring forest patch). If so, the overall benefits from pollination within a given landscape are expected to peak at intermediate levels of natural/semi-natural habitat cover, at the point where the spatial interface with crops is maximised (Morandin & Winston 2006; Figure 14d). The spatial arrangement of the habitat patches within the landscape is likely to matter; modelling (Keitt, in review) suggests that pollination benefits are optimised for maximized by providing islands of nesting habitat where inter-island distance matches mean foraging and dispersal distances of wild pollinators. Hence, strategically creating new patches of habitat in areas of intensive agriculture (for the crops that benefit from wild pollination) can increase services from wild pollination significantly. The size of these islands is likely to be important, though, as pollinator communities may become unviable in very small fragments (Keanrs et al. 1998). For some pollinator species, agricultural areas (the crop fields themselves) are also important habitats, providing nesting and floral resources. In this case, even small natural or semi-natural habitats (too small to support pollinator populations by themselves), can bolster populations of important pollinators. 4.1.4 What are the main threats to the provisioning of wild crop pollination? Loss of suitable habitat is recognised as a key driver of declines in wild pollination (e.g. Steffan-Dewenter 2003; Ricketts et al. 2004; Morandin & Winston 2006; Priess et al. 2007). Habitat degradation, for example through agricultural intensification, leads to scarcity in critical floral and nesting resources for many species. The use of chemicals (such as insecticides, herbicides and fertilisers) in conventional agriculture has been found to reduce populations of pollinators (Kearns et al. 1998). Climate change is emerging as a potentially new threat, as phenological shifts may result in the disruption of plant-pollinator interactions (Memmott et al. 2007). Invasive non-native species (including plant, mammals and insects) pose additional threats (Cole et al. 1992; Kearns et at. 1998). 4.1.5 Are abrupt changes likely in the provision of wild crop pollination? It is possible that a threshold in pollinator species/functional diversity, exists below which pollination services become too scarce or too unstable (Klein et al. 2007). Such a tipping point might occur when, at a landscape context, sufficient habitat is destroyed that the next 32 marginal change causes a population crash in multiple pollinators. Modelling supports this prediction (Morandin & Winston 2006; also Keitt in review). Alternatively, a threshold in habitat loss may lead to the collapse of particularly important pollinators, leading to a broader collapse in pollination services (pollinator keystones). Supporting this prediction, Larsen et al. (2005) found that large-bodied pollinators tended to be both most extinction-prone and most functionally efficient, contributing to rapid functional loss with habitat loss. Empirical data on changes in the provision of pollination services to agriculture are still very sparse. However, studies in California clearly indicate that there is indeed a tipping point in the provision of such services from wild pollinators (Kremen et al. 2002, 2004; Larsen et al. 2005: Greenleaf & Kremen 2006; Klein et al in prep; Chaplin-Kramer et al. in prep). From this work we can extrapolate that in most areas of California’s agricultural region such threshold has already been surpassed and there is now little wild pollination. This makes crop production substantially to entirely reliant on managed honey bees, whose numbers show strong within-year variation and an overall declining trend due to diseases and other factors (NRC 2006). The studies in California are illuminating because the region has very strong gradients, ranging from small farms surrounded by natural habitat to some of the most intensively managed agricultural landscapes in the world. Given global trends in agricultural intensification, the California region is therefore very informative of the possible future changes in the provision of pollination services worldwide. A model currently being developed by the NCEAS working group could be used in the future to explore such thresholds. Overall, we predict that there is a medium to high probability that the provisioning of wild pollination services is likely to be subject to thresholds/tipping points in the foreseeable future (by 2025), with a very high probability that such thresholds will happen in regions of very intensively managed agriculture. What this will mean for actual crop productivity is less clear as there are substitutes for wild pollinators and for pollinator-dependent crops. 4.1.6 Can we quantify and map the global provision of this wild crop pollination and how it might change? State of knowledge and data availability Wild pollination is a valuable ecosystem process because it can increase the yield of economically important crops. Phase II of the Review on the Economics of Biodiversity Loss will attempt to assess how such services are affected by biodiversity loss (e.g. by deforestation leading to declines in habitat for wild pollinators). An ideal answer to this would involve producing a global map of pollination services (measured in units reflecting the contribution to increased yields per ha per year). Such a map would be generated for different scenarios of possible global changes (e.g. what if all remaining Andean cloud forest is lost?), and differences in pollination services would be contrasted to evaluate the extent to which pollination is affected by those changes. The economic valuation would be 53 done on top of this ‘biophysical model’, accounting for local/regional variation in crop yields as well as for global aspects such as changes in commodity market prices. Although a biophysical model of delivery of pollination services is not fully in use as yet, two major efforts are currently being reconciled with the aim to have such a model in operation by October 2008: e Kremen et al. (2007) have created a conceptual model of ‘mobile agent-based ecosystem services’, which forms the basis of work carried out by the NCEAS Working Group on pollination services. The model includes interactions and feedbacks among policies affecting land use, market forces, and the biology of organisms involved. The group is conducting quantitative syntheses of the key relationships within this model, which has already been done for the dependence of world crops on pollinators by Klein et al. (2007) and for the relationship between distance from natural habitat and pollination services by Ricketts et al. (2008), and is underway for impacts of disturbance on bee abundance and diversity (led by Rae Winfree), and pollinator functional traits (led by Neal Williams). This component considers the sensitivity of pollinator species based on life history and other traits (e.g., body size, trophic specialization, nesting habit) and explores the consistency of resulting functional group responses among different types of disturbance. From this, a quantitative model will be built. e The Natural Capital Project is developing tools for modelling and mapping the delivery, distribution, and economic value of ecosystem services and biodiversity. The Project’s pollination module is developing a biophysical model for pollinator abundance on crops in a landscape. It uses information on key pollinators, the availability and location of their nesting and floral resources on the landscape, and their flight ranges, to predict their abundances on crops. The result is a map of relative pollinator abundance in agricultural fields or pixels. Predicted effects on yield or quality of crops have not yet been calculated, but the group plan to do so. Recently, these two efforts have merged, with key leaders of the NCEAS group helping to devise the Natural Capital model. By October 2008, they plan to have the model formulated and validated on three landscapes by comparing model predictions against empirical data. In a related study, Priess and colleagues (2007) used empirical relationships on the effects of forest distance on both pollinator diversity and fruit set of coffee (based on the results of field experiments carried out by Klein et al. 2003a,b,c) to estimate future changes in pollination services for different land use scenarios in Sulawesi, Indonesia. These efforts indicate that it is possible to develop a landscape-scale spatially-explicit model based on an empirical understanding of how different crops benefit from wild pollination (Klein et al. 2007) and how pollination decays with distance from natural/semi- natural habitat (Ricketts et al. 2008), ideally complemented with more detailed data on the identity/guild of the pollinators, the specific pollinator needs of crops, and the value of modified habitats (including the agricultural lands themselves) for nesting and floral resources. Currently available data are far from perfect, but allow for broad generalisations, and expert opinion could be used to fill some data gaps. The key challenge would reside in extrapolating these models for application at a global scale. Indeed, global maps would be needed not only on the distribution of the relevant 54 natural/semi-natural habitats, but also on the distribution of the relevant crops (that benefit from pollination). Furthermore, these maps would need to have sufficient spatial detail to be relevant to the fine-scale spatial dynamics that governs pollination, as foraging ranges of pollinators have been empirically shown to be typically less than 1 km. Such maps currently do not exist. The main limiting factor is probably the unavailability of a detailed map of global crops distribution, with information for individual crops and with sufficiently fine spatial resolution. But the production of fine scale maps of natural/semi-natural habitat (with information on their relative value for pollination services) is also not trivial. Such maps can be generated for particular regions (e.g. Natural Capital model now being applied to the Valuing the Arc project in Tanzania) but certainly not for the world within the time frame of the Review. Two possibilities exist in the short-term for the development of a first-cut pollination model at the global scale: e A probabilistic approach, whereby the probability of each pixel being under a crop is estimated from data on the pixel’s suitability for crops (Fisher et al. 2002) as well as broader crop maps (Cassman et al. 2005; Ramankutty et al. 2008). The expected benefits from pollination on each pixel would then be estimated using a generalised biophysical model. e A sampling approach, in which detailed models are developed for each of a set of representative landscapes, applied using fine-scale maps of crops and habitats for those landscapes. Expected global benefits from pollination, and predicted changes for different scenarios, would then be obtained by extrapolation using global crop data (Cassman et al. 2005; Ramankutty et al. 2008). We believe the second approach is the most feasible as well as the most informative.The probabilistic approach would allow for the consideration of uncertainty of what the relevant crops are in each area, but the spatial resolution would not be sufficiently fine for the adequate incorporation of the relevant spatial dynamics. The sampling approach would benefit from finer spatial data on the distribution of both crops and natural/semi-natural habitats, therefore resulting in substantially more reliable predictions. A choice of adequate sample landscapes (including temperate and tropical regions, containing strong gradients in agricultural intensification) would allow for a first-cut extrapolation to the global scale. We therefore recommend that a global evaluation of the effects of biodiversity loss and ecosystem change on pollination services builds from the work currently being done by the Natural Capital Project, which is focusing on four landscapes: California, Tanzania. China and Hawaii. Even at the finer landscape scale, it is important to keep in mind the limitations of any model that can possibly be developed and applied in the near future. Smaller plots of agricultural land in close proximity to natural biodiversity, which are important for pollination services, are unfortunately likely to be lost due to coarse resolution of the data. Small field sizes and mixed cropping systems make Africa particularly vulnerable to this. A model in the lines currently being developed and proposed here would be particularly informative for the effects of changes in wild pollination services associated with changes in land cover, an less useful to predict variation in other factors. Effects of variation in pesticide use can somehow be incorporated in the classification of crop types (as intensive crops tend to make more use of pesticides) but only up to a certain extent (as, within any given crop type, level of pesticide use depends on individual farmer decisions). The model 55 would be unlikely to capture changes dues to invasive species, unless the value of invasive floral resources for pollinators is know along with habitat associations of these plant species (one example for which this could be known is Centaurea solstitialis, yellow star thistle). Adequacy of scenarios A probabilistic approach would require the generation of predicted global maps of land cover, including predictions for the distributions of particular sets of crops, at a fine spatial scale. Approaches such as those developed by the ATEAM Project (Advanced Terrestrial Ecosystem Analysis and Modelling) for Europe (Schréter et al. 2004; Rounsevell et al. 2006) could potentially be expanded to the global scale (within 1 year?) A sampling approach would require the generation of detailed maps of predicted land cover for a set of representative regions, for each of the chosen scenarios. While these are not trivial to generate either, results would be more realistic and a finer scale than for when produced at a global scale. Approaches such as those used by Priess et al. (2007) seem to have potential. They used ‘generalized cellular automata’ to allocate future land use based on biophysical suitability (climate, soils, topography), allocation factors (e.g. distance to the next river, land use on adjacent cells, preferred walking distance to field), demography (population growth rate, migration), and the land use strategies of farmers (e.g. moderate or high intensity agriculture, forest use) and other restrictions like the protection status of the area. The extrapolation from study landscapes to the global scale would require linking global coarse scenarios for changes in land use distribution with the finer landscape scenarios. This could potentially draw on the methods and approaches that have been developed for downscaling land use scenarios (e.g. Verburg et al. 2006). Model validation is fundamental, by matching modelled predictions to exisiting data from targeted areas. A potential approach would be to select existing data sets, while another would be to prioritize study of pollination service at particular target areas for which other data are known. A fundamental variable in the scenarios is the extent to which managed pollinators (particularly honeybees) are made available in crops. The scenarios being contrasted would need to consider that this is the same in both cases. Possible extreme scenarios include, on the one hand, a situation in which wild pollination becomes the only available mechanism for crops that depend on biotic pollination (in the remote possibility that Colony Collapse Disorder eliminates all managed honeybees) and, on the other hand, a situation in which managed pollination is provided whenever useful and the value of wild pollination is simply by complementing (or interacting with) managed pollinators. Reality is somewhere in between, and so scenarios could also consider an intermediate situation in which particular crops, from particular regions, benefit from managed pollination and others do not. The key consideration would be that any pair of scenarios being compared would need to consider matching levels of managed pollination. The same applies for levels of pesticide use. 56 4.1.7 Insights for economic valuation Appropriate scenarios can generate maps of different states of the world, varying in their spatial patterns of distribution of crops and natural/semi-natural habitat. A model along the lines of those described above can then be used to create surfaces of pollination services (measured in changes in the yields of particular crops) for each state of the world. The contrast between these two maps will indicate areas where crop yields are expected to change (either increase or decrease). Data on the market value of these crops (adjusted for possible variations in global crop prices arising from differences in the scenarios) could then be used to monetise the map of variation in crop yields. 4.1.8 Some key resources The National Center for Ecological Analysis and Synthesis (NCEAS) Working Group on “Restoring an ecosystem service to degraded landscapes: native bees and crop pollination” is synthesising and analysing data on bee populations, pollinator communities and pollination services across agro-natural landscapes, including the development of models for the management of pollination services. The Group is led by Claire Kremen and Neal Williams and composed of c. 20 experts, including amongst others Alexandra-Maria Klein, Ingolf Steffan-Dewenter, Rachael Winfree and Taylor Ricketts. The Natural Capital Project, a joint venture among The Woods Institute for the Environment at Stanford University, The Nature Conservancy, and the World Wildlife Fund, is developing tools for modelling and mapping the delivery, distribution, and economic value of ecosystem services and biodiversity. It includes a pollination module (led by Taylor Ricketts), which is developing a biophysical model for pollinator abundance on crops in a landscape. Modelling for the NCEAS Working Group and the Natural Capital pollination module is being led by Eric Lonsdorf. ' This is our recommendation for Phase II, based on the results of this review. It does not commit the leaders of Phase II to follow it, and it does not commit the recommended research group to actually do such work. 57 e The EU-funded ALARM (Assessing Large-scale Risks for Biodiversity with Tested Methods) project, led by Ingolf Steffan-Dewenter, includes a research module on pollination investigating the risks resulting from, and rates and extent of, loss of pollinators. With a particular emphasis on Europe, this 5-year (2004-2009) project is developing predictive models for pollinator loss and consequent risks. e The Millennium Assessment (Cassman et al. 2005) used data on rain-fed and irrigation croplands obtained from a cropland-focused reinterpretation of GLCCDv2, based on methods described in Wood et al. (2000). e Ramankutty et al. (2008) developed a new global data set of croplands and pastures circa 2000, at 5 min (~10 km) resolution, by combining agricultural inventory data and satellite-derived (MODIS-derived and GLC2000) land cover data. e The Global Agro-Ecological Zones (GAEZ) assessment created a global database of estimated yields for various crops by matching the soil type, terrain, and climate of grid cells with productivity levels for 154 land-use types documented in places with similar characteristics (Fisher et al. 2002). The 5-arc-minute grid generated by the GAEZ assessment contains measures of agricultural suitability, assigned to eight categories based on estimated productivity levels as a percentage of maximum observed yields for a given crop. e The International Pollinators Initiative, established by the Fifth Conference of Parties to the Convention on Biological Diversity in 2000, declared an ‘urgent need to address the issue of worldwide decline of pollinator diversity’. 4.1.9 Participants Authors James J. J. Waters, Ana S. L. Rodrigues, Andrew Balmford (University of Cambridge, UK) Taylor H. Ricketts (WWF-US, US) Claire Kremen, Alexandra-Maria Klein (University of California, Berkley, US) Neal M. Williams (Bryn Mawr College, US) Contributors Saul A. Cunningham (CSIRO, Australia) 58 4.2 Biological control in crops This section is based on a quick literature review, and did not receive contributions from, nor has it been reviewed by, experts in this field. 4.2.1 Why is biological control in crops important for human wellbeing? Biological control is the process by which one organism reduces the population density of another organism, for example through predation or parasitism. In nature, most organisms are consumed by other organisms, which in many cases leads to drastic reductions in the population of the prey species; in biological control, man exploits this ‘natural control’ to suppress the numbers of pest species (Bale et al. 2008). Here we focus on the value of biological control for reducing the effects of pests in agricultural systems, and therefore contributing for increases (or preventing decreases) in crop yields. We recommend a focus on biological control for crops mainly because of the availability of data (e.g. Losey & Vaughan 2006) and the huge value of crop plants to human well-being. Agricultural pests cause significant economic losses worldwide. Globally, more than 40% of food production is being lost to insect pests, plant pathogens, and weeds, despite the application of more than 3 billion kilograms of pesticides to crops, plus other means of control (Pimentel 2008). In the US alone, it is estimated that more than US$18 billion are lost due to insect damage (including more than US$ 3 billion spent in insecticides), of which about 40% ate attributed to native species and the remaining to exotic pests (Losey & Vaughan 2006). These values, however, would be much higher if biological control was not in place. Losey & Vaughan (2006) estimate that 65% of potential pest species are being suppressed in the US, with a total value of pest control by native ecosystems around US$ 13.60 billion. Through a predator removal experiment, Ostman et al. (2003) showed that the presence of natural enemies increased barley yields 303 kg/ha, preventing 52% of yield loss due to aphids. Biological control in crops may be totally natural, without direct intervention from man, or it may be enhanced through biological control interventions. Frequently, the term biological control refers to the latter only, when an organism is used by man to reduce the population density of another organism (Bale et al. 2008). Here, however, we are interested in the contribution of ‘wild nature’ into both of these, and particularly in the natural control that flows from unmanaged ecosystems. Natural control of plant pests is provided by generalist and specialist predators and parasitoids, including birds, bats, spiders, ladybugs, mantis, flies, and wasps, as well as entomopathogenic fungi. In the short-term, this process suppresses pest damage and improves yields, while in the long-term maintains an ecological equilibrium that prevents herbivore insects from reaching pest status (Zhang et al. 2007 and references therein). In biological control interventions, pest control is enhanced through a diversity of approaches (Bale et al. 2008 and references therein): - Inoculative control is used mainly against ‘exotic’ pests that have become established in new countries or regions of the world and involves the collection and 59 release of natural enemies from the country or region of origin of the pest. This works better with perennial crops (fruit plantations and forests), where the long- term nature of the ecosystem enables the interactions between pest and natural enemy to become fully established over a period of time. - Augmentative control involves periodically reintroducing natural enemies that are usually commercially produced for that purpose. It is particularly indicated for short-term annual crops, as populations of natural enemy species may not persist between the crop production cycles. This includes both ‘inundation’ (the mass production and release of large numbers of the control agent) and ‘seasonal augmentation’ (where the natural enemy population is built-up through successive releases during the same growing season). - Conservation control (see Landis et al. 2000 for a review) refers to the manipulation of the environment to increase the effectiveness of indigenous predators and parasitoids, usually against native pests. Various measures are implemented to enhance the abundance or activity of the natural enemies, including manipulation of the crop microclimate, creation of overwintering refuges (like ‘beetle banks’), increasing the availability of alternative hosts and prey, and providing essential food resources such as flowers for adult parasitoids and aphidophagous hoverflies. Biological control reduces rather than eradicates pests, such that the pest and natural enemy remain in the agro-ecosystem at low densities. A number of important pests can be kept at a low population density by biological control over long periods of time. In other cases, populations of pests are significantly reduced by natural enemies, but repeated additional methods are needed to achieve an adequate level of control (e.g., resistant plants, cultural techniques, physical barriers, semiochemicals and, as a last resort, the use of selective chemicals) — this is the fundamental philosophy of integrated pest management (Bale et al. 2008). Here we are particularly interested in the value wild nature to natural (unmanaged) pest control and to (managed) conservation control. However, wild nature also contributes to inoculative and augmentative pest control through the provision of natural enemies in the regions of origin of exotic pests. 4.2.2 What are the overall trends in the biological control of crops? The Millennium Assessment indicated a medium to high certainty that pest regulation services are declining (MEA et al. 2005). Agricultural intensification (reduction in the landscape complexity, increase in pesticide use) and loss of crop genetic diversity increase the exposure of crops to major disease outbreaks and therefore the importance of biological control. At the same time, improvement in artificial pest control (through the use of pesticides) and the development of pest-resistant crops reduce the need for biological control. However, pesticide use has an associated array of problems (Pimentel 2008): they are expensive and need to be applied frequently; public health and environmental impacts, with more than 26 million people are poisoned annually from pesticides, resulting in more than 220,000 deaths; and environmental impacts such as losses to non-target species, including natural enemies of crop pests. Indeed, pesticides can be more damage to natural enemy population 60 than to pests, therefore exacerbating the very problem they were designed to solve (Kremen & Chaplin-Kramer 2007). Furthermore, pesticide-resistant organisms evolve rapidly (e.g. insects usually evolve resistance within about a decade of the introduction of an insecticide; Palumbi 2001), while there are no known cases of evolved resistance of natural enemies (Bale et al. 2008). The value of natural control should therefore not be underestimated or presumed to be declining, even less so as demand for organic products is increasing (Willer & Yussefi 2008). Data on the populations of biological control agents are scarce but the trends are presumed to be negative due to habitat transformation associated with agricultural intensification (agricultural expansion, enlargement of field size, removal of non-crop habitat resulting in a loss of natural landscape features required for maintaining their populations) and increase in pesticide use. On the other hand, the increase in organic farming worldwide may help to reverse this trend (Bengtsson et al. 2005; Willer et al. 2008). 4.2.3 How is the provision of biological control in crops affected by changes in wild nature? Biological control seems to be greatly determined by the abundance of natural enemies, and indeed this is often used as a measure of the intensity of biological control (e.g. studies reviewed in Bianchi et al. 2006 and Kremen & Chaplin-Kramer 2007). Several studies demonstrated increases in predation rate with predator abundance, while for parasitoids the majority of evidence for increased densities affecting pest control comes from exotic introduction examples (see references in Kremen & Chaplin-Kramer 2007). However, enhanced enemy abundance or richness does not necessarily guarantee improved pest control (see below). The diversity of natural enemies also seems to improve biological control, through a variety of mechanisms, including: species complementarity, when more than one type of predator or parasitoid adds to the control of a pest species; sampling effect, meaning a particularly effective natural enemy is more likely by chance alone to occur when more species are present; redundancy, having more species will buffer against disturbance or ecosystem change; and idiosyncrasy, when the whole is greater than the sum of the parts due to interaction between species (Tscharntke et al. 2005; and see Kremen & Chaplin-Kramer 2007 for a review). However, these expected benefits of diversity to pest control may be countered by antagonistic effects between different natural enemy species, such as competition and intraguild predation (Kremen & Chaplin-Kramer 2007). On the other hand, intraguild predation could help supplement natural enemy diets when pest densities are low, potentially maintaining a higher natural enemy population size more capable of controlling sudden pest outbreaks. The overall effect seems to be context-dependent, being important to consider both the diversity and the composition of natural enemy assemblages (Wilby & Thomas 2002; Kremen 2005), as well as the characteristics and ecology of predators, and the identity of their prey species (Kean et al. 2003; Wilby et al. 2005). While species richness will be most important for the control of some pests, the presence of a particular predator or parasite will be of utmost importance for others (Wilby & Thomas 2002). The provision of biological control is crucially determined by the identity of biological control organisms, which may even have a stronger effect than diversity (e.g. Straub & Snyder 2006). Non-crop habitats are fundamental for providing habitat where many biological control agents (predators, parasitoids) mate, reproduce, and overwinter (Zhang et al. 2007) as well 61 as additional, complementary, food (Kremen & Chaplin-Kramer 2007). The quality of the habitat can make a significant difference (e.g. Thies et al. 2005; Tylianakis et al. 2007) and this can be improved through habitat management (Landis et al. 2000). Landscape diversity or complexity is generally positively associated with the abundance and species richness of natural enemies and may be the most crucial factor driving biological control services (for reviews see Bianchi et al. 2006; Kremen & Chaplin-Kramer 2007; Tscharntke et al. 2007). Landscape complexity can be measured by a variety of factors, including the proportional area of natural habitat, distance to natural habitat or perimeter to area ratio of the field (Kremen & Chaplin-Kramer 2007), and beta diversity (species turnover) across natural habitat patches (Tscharntke et al. 1998). Landscape connectivity (Tscharntke et al. 2007) and age of natural habitat (Tscharntke et al. 1998) also seem to increase biological control. While there is evidence that increased landscape complexity is correlated with diversity and abundance of natural enemy populations, this does not necessarily translate into improved pest control, as pest densities may also respond positively to landscape complexity (Thies et al. 2005). For example, while 75% of the studies in Bianchi et al. (2006) found increased abundance of natural enemies in more complex landscapes, only 45% showed lower pest pressure. Relatively very few studies measured actual predation rates or yields through controlled experiments (e.g. Ostman et al. 2003; Williams-Guillen et al. 2008). The proximity to semi-natural habitats greatly influences biological control by increasing natural enemy abundance, diversity, predator/prey ratios and importantly predation risk or parasitism for crop pests (Tscharntke et al. 2007; Bianchi et al. 2006). The distance of enhanced parasitism varies from tens to hundreds of metres, depending on parasitoid species (Tscharntke et al. 2005). Although there are many examples of parasitism rates being higher at the edges of fields than the middle, they are also frequently found to be unrelated, so that one cannot be sure whether the number of natural enemies or their movement is limiting biological control (Tscharntke et al. 2007). Unfortunately, there are few experimental studies (e.g. Tscharntke et al. 1998) quantifying the relationship between distance from semi-natural habitat and biological control, and no work has investigated the shape of a general relationship, as has been done for pollination (Ricketts et al. 2008). Habitat fragmentation may also affect biological control (With et al. 2002), with a greater effect on specialist parasitoids than generalist predators (Tscharntke et al. 2007), such that generalist and mobile enemies may even profit from the high primary productivity of crops at the landscape scale, their abundance compensating for losses in diversity. Natural habitat may actually have negative effects (“ecosystem disservices’) on biological control, particularly as a source of pests (Thies et al. 2005; Kremen & Chaplin-Kremer 2007). On the other hand, some studies have shown a greater effect of habitat loss on the predators or parasitoids than on their herbivore prey and so the disservice of pest provision may be negated by the benefits of natural enemies in large and well-connected habitats (Kremen & Chaplin-Kremer 2007 and references therein). Crop diversity can reduce the incidence of pests and diseases, both the increase in the genetic diversity of individual crops (e.g. Zhu et al. 2000) and the increase in the diversity of crop species in multi- or poly-cropping systems (e.g. Sutherland & Samu 2000). 62 Relationship between habitat area and biological control Work quantifying the effect of the area of natural habitat on the provision of natural enemies and biological control services is lacking (Kremen & Chaplin-Kramer 2007), although results suggest that may be greater for parasitoids/predators than their prey (e.g. Kruess & Tscharntke 2000). The question is complicated by the different dispersal abilities of insect species: while a large uncultivated area may provide a large source of natural enemy individuals and species, the same area distributed in smaller fragments among cultivated fields would allow greater dispersal (Tscharntke et al. 2007). The optimal resolution to this tradeoff will also depend on the attributes of the pests and the crops, as ephemeral annual crops will benefit most from natural enemies moving quickly into the area (Bianchi et al. 2006) whilst a larger refuge of diverse natural enemies might be more suitable for perennial crops. Equally, highly mobile pests may be best controlled by natural enemies in a landscape of well-connected natural habitats (Tscharntke et al. 2007). 4.2.4 What are the main threats to the provision of biological control in crops? The main threat to the provision of biological control seems to be habitat loss and degradation, particularly associated with agricultural intensification. This results in landscape simplification, and natural habitat fragmentation leading to losses in the diversity and abundance of natural enemies. Intensive agricultural management practices threaten natural enemy populations (Naylor & Ehrlich 1997). Ironically, pesticide use has had severe detrimental effects on natural enemies, often greater than on the target pests, and may even result in the emergence of new pests (Kremen & Chaplin-Kramer 2007). Natural enemy species will also be under threat from invasion of foreign species and the disruptive effects of climate change. 4.2.5 Are abrupt changes likely in the provision of this benefit/process? The relationship between densities of natural enemies and the biological control services they provide is unlikely to be linear (Losey & Vaughan 2006). From the complexity of interactions and non-linearity of the relationship between diversity and function mentioned above, it seems likely that a function may decline disproportionately when a tipping point in natural enemy diversity is passed. However, there does not seem to be empirical evidence to prove this. Thresholds in landscape structure (interpatch distance) have been shown to impact the aggregative response of predators, thereby generating a similar threshold in pest populations (With et al. 2002). The importance of natural enemy assemblage composition in some instances of biological control (Shennan 2008) indicates that changes in composition can lead to disproportionately large, irreversible and often negative shifts in ecosystem services (Diaz et al. 2005). Overall, we predict that there is a medium to high probability that the provisioning of biological control is subject to thresholds/tipping points in the foreseeable future (by 2025), with a very high probability that such thresholds will happen in regions of very intensively managed agriculture. 63 4.2.6 Can we quantify and map the global provision of biological control in crops, and how it might change? Making assumptions on the proportion of insects that cause any damage, the proportion of native pest species in the US, and the relative amount of pest suppression that is due to insects rather than other causes, Losey and Vaughan (2006) illustrate how to use the ratio of cost of damage due insect pests to estimate the value of native pest control services. However, such a calculation is based on location-specific assumptions, and it is not clear that it could be scaled up to a global calculation. It would also not be fine enough to quantify the effects of changes in biological control associated with changes in states of the world due to specific policy measures. A model for biological control services has not been created, although Kremen et al. (2007) created a conceptual framework for the effects of land-use change on ‘mobile-agent-based ecosystem services’, including pest control. Initially developed for pollination (see chapter), the model includes interactions and feedbacks among policies affecting land use, market forces, and the biology of organisms involved. Unlike for pollination (section 4.1), syntheses of the literature to quantify the key relationships within this framework have not yet been carried out for biological control. These include: - The relative value of biclogical control to crop productivity, that is, their vulnerability to a shortage of natural enemies. This can be identified by the intensity of pesticide applications to those crops (Kremen & Chaplin-Kramer 2007). - The relationship between the provision of biological control and distance to natural habitat patch (related to the dispersal and foraging movements of natural enemies). - The relationship between the provision of biological control and area of habitat patch. - The functional significance of species diversity in each system would also be a requisite for determining natural enemy assemblage service provision. These syntheses need to build from field data and (ideally) experiments. We are not sure if the current body of knowledge is sufficient for such syntheses to be produced from published information, or if additional field work would be required. A complication in relation to pollination (in which natural habitat is always expected to have a positive effect), is that natural habitat may simultaneously result in the increase in the populations of natural enemies and of the pests. Hence, the measure of enemy abundance on its own is not an adequate surrogate of biological control and instead information on the variation in crop yields should be used. Adequacy of scenarios We suspect that the same variables required for modelling pollination (section 4.1) would be needed for modelling biological control. 64 4.2.7 Insights for economic valuation See pollination section (section 4.1) 4.2.8 Some key resources This section was based on a (quick) overview of the literature, and we suggest it is further developed by contacting experts in this field. We recommend the following as first points of contact: e Claire Kremen at the Department of Environmental Science, Policy, and Management, University of California, Berkeley (USA) who has led development of a model on ‘mobile-agent-based ecosystem services’, including pest control (Kremen et al. 2007). e Teja Tscharntke in the Department of Crop Science, University of Gottingen, (Germany), whose group also includes Carsten Thies. e David Pimentel at the Department of Entomology, College of Agriculture and Life Sciences, Cornell University (USA). e Felix Bianchi at the Functional and Spatial Ecology group in CSIRO Entomology (Australia). 4.2.9 Participants Authors James Waters, Andrew Balmford and Ana Rodrigues (University of Cambridge, UK) 65 4.3 Genetic diversity of crops and livestock This section is based on a quick literature review, and did not receive contributions from, nor has it been reviewed by, experts in this field. 4.3.1 Why is the genetic diversity of crops and livestock important for human wellbeing? Food production and security depend on the wise use and conservation of agricultural and livestock biodiversity and genetic resources. Crops and their wild relatives (plant genetic resources for food and agriculture; FAO 1997) and livestock and their wild relatives (animal genetic resources for food and agriculture; FAO 2007) have the genetic variability that provides the raw material for breeding new crop varieties, through classical breeding and biotechnological techniques. This diversity is fundamental to respond to environmental and demographic changes (Esquinas-Alcazar 2005). While thousands of plant and animal species have been domesticated for human use, currently, barely more than 150 plant species are cultivated and most of mankind lives off no more than 12 plant species (Esquinas-Alcazar 2005) and 30 crops supplying 90% of the global calorie intake (Wood et al. 2005). Domestic animals provide 30% of food globally but just 14 species represent 90% of the global production (Wood et al. 2005). http://www.nature.com/nrg/journal/v6/n12/full/nrg1729.html - B3Global population growth and corresponding rising demand for food favoured the adoption of highly productive varieties. The introduction of modern farm machinery, marketing and transport methods that require uniform crop and livestock characteristics have required the introduction of standard, homogeneous plants and animals. This trend, which peaked during the so-called 'Green Revolution’, has made it possible to boost food production, but lead to the loss of innumerable heterogeneous (and often locally adapted) traditional farmers’ varieties (Esquinas-Alcazar 2005). The loss of local species and varieties usually results in irreversible loss of the genetic diversity they contain, known as genetic erosion. This has dangerously shrunk the genetic pool that is available for natural selection, and for selection by farmers and plant and livestock breeders, and has consequently increased the vulnerability of agricultural crops and livestock production to sudden changes in climate, and to the appearance of new pests and diseases (Esquinas-Alcazar 2005). For example, in the United States in 1970, the fungus Helminthosporium maydis destroyed more than half the standing maize crop in the southern part of the country. The crop had been grown from seeds that have a narrow genetic base and are susceptible to this disease. In this case and others, the problem was resolved by breeding resistant varieties using genetic resources that were obtained from other parts of the world (Esquinas-Alcazar 2005). The conservation of animal and plant genetic resources for food and agriculture relies on the preservation of both the variety of domesticated species and their wild relatives. This can take place ex-situ, particularly for plants, through the development of seed banks and germplasm collections (FAO 1997; e.g. the recently inaugurated Svalbard Global Seed Vault, already with 100 million seeds). Here, we are particularly interested in the in-situ conservation of genetic diversity, through the protection of the natural habitats of wild relatives or through the maintenance of traditional agricultural landscapes. Note that with 66 the exception of the wild boar (Sus scrofa) the ancestors and wild relatives of major livestock species are either extinct or highly endangered and that therefore domestic livestock are the only depositories of the now largely vanished diversity of the wild ancestors. This is a major difference from crop species, in many of which the wild ancestors are commonly found at the centres of origin and represent an important source of variation and adaptive traits for future breeding programmes (FAO 2007). The term ‘crop biodiversity’ is sometimes used to refer to the diversity of crops species within one region (e.g. wheat and barley rather than just wheat; e.g. Di Falco & Perrings 2005). We are not addressing the effects of crop diversity in that sense. Despite the widespread use of modern varieties, landrace (locally adapted, traditional varieties) are still grown in many parts of the world. This is particularly the case in marginal areas where locally-adapted varieties (e.g., to specific soil and water regimes) thrive better. Landraces may also allow labour requirements to be spread (through differences in maturing time of the varieties), be adequate for multiple usages (e.g., differences in stalks can make them better for use as fencing material or for feeding to cattle), and reduce disease/pest risk (Cassman et al. 2005). 4.3.2 What are the overall trends in the genetic diversity of crops and livestock? The genetic diversity of crops and livestock has been declining markedly, but the overall rates of loss are not easy to quantify (FAO 1997). For example, in China nearly 10,000 wheat varieties were in use in 1949, but only 1,000 were still in use by the 1970s. In the US, 86% of the apple varieties, 95 % of the cabbage, 91 % of the field maize, 94 % of the pea, and 81 % of the tomato varieties documented as having been in use between 1804 and 1904 apparently no longer exist (FAO 1997). Of the 7,616 livestock breeds listed by FAO’s Global Databank for Animal Genetic Resources for Food and Agriculture, around 20 % are classified as at risk and 62 breeds became extinct during the 2001-2007 period — amounting to the loss of almost one breed per month (FAO 2007). The three main crops — rice, maize and wheat — that provide over half the world population’s requirement for protein and calories (Bioversity International 2008) are increasing reliant on a small number of modern varieties. Indeed 80% of the wheat area sown in developing countries is modern semi-dwarf varieties and over 75% of all rice in Asia is improved semi-dwarf varieties (Cassman et al. 2005) 4.3.3 How is the genetic diversity of crops and livestock affected by changes in wild nature? The genetic diversity amongst wild relatives of crop (and to a lesser extent livestock) species is a direct result of the species and population diversity amongst these species. They are therefore affected by changes that lead to species/populations extinctions, such as loss and degradation of natural habitats. Amongst domesticated relatives, loss of varieties and diversity within varieties is expected from the loss of area under traditional agricultural practices and associated local breeds, such as when these are either abandoned or converted to intensive agriculture. 67 Relationship between habitat area and genetic diversity of crops and livestock Losses in diversity species with area are likely to follow a non-linear relationship (linear with log area) similar in shape to the species-area relationship (but probably much steeper for the loss of intra-specific/intra-variety genetic diversity). 4.3.4 What are the main threats to the genetic diversity of crops and livestock? The main threat to the genetic diversity of crops and livestock is the marginalization of traditional production systems and the associated local breeds, driven mainly by the rapid spread of intensive agriculture and livestock production, often large-scale and utilizing a narrow range of crop varieties and livestock breeds (FAO 1997, 2007). The intensification process has been driven by rising demand for agricultural and animal products, and has been facilitated by the ease with which genetic material, production technologies and inputs can now be moved around the world (Esquinas-Alcazar 2005; FAO 2007). Acute threats such as major disease epidemics and disasters of various kinds (droughts, floods, military conflicts, etc.) are also a concern — particularly in the case of small, geographically concentrated varieties and breed populations (FAO 2007). The loss of cultural diversity is also a factor in the loss of genetic diversity as cultural knowledge about production of traditional varieties and extraction of wild relatives is also lost (FAO 1997). Wild varieties are threatened by habitat _loss_ and degradation, and overharvesting. We presume that climate change may also affect not only wild relatives but also domestic breeds and varieties, if it reduces the degree to which they are adapted to their local environments. 4.3.5 Are abrupt changes likely in the genetic diversity of crops and livestock? Given the likely non-linear relationship between area and genetic diversity, we predict that in some cases a small change in area (of natural habitat, or of traditional agricultural lands) may result in a disproportionate loss in genetic diversity of crops and/or livestock. This is probably more likely in areas that have already suffered extensive habitat loss and land conversion, where the remaining populations of particular varieties and breeds are quite small. Climate change may also have non-linear effects on genetic diversity of crops and livestock. - Overall, we suspect that abrupt changes are likely in the genetic diversity of crops and livestock in the foreseeable future (by 2025), particularly in regions subject to extensive habitat loss and land conversion. 68 4.3.6 Can we quantify and map the global diversity of crops and livestock, and how it might change? We predict that it would be possible to create a global, spatially-explicit, model of crop and livestock diversity through the following steps: I) 2) 3) Map the distribution of wild relatives and of local varieties and breeds. There seem to be a few databases on which this could be based, including for example the Domestic Animal Genetic Resources Information System (DAGRIS), and the databases under the Germplasm Resources Information Network (GRIN). However, we are not sure if these are geographically and taxonomically comprehensive and if the distributional information (which seems to be at the country level) is sufficient. Obtain an estimate of the economic value of each breed/variety as insurance against pest and disease outbreaks. This would need to be estimated from statistics on past outbreaks to predict the risk of future outbreaks. The economic consequences of those outbreaks would be estimated from the predictions of damage and of crop/livestock prices. The economic value of each breed/variety would be calculated as the predicted avoided damage by using its genetic diversity to prevent outbreaks. This would in principle follow similar principles to the calculation of an insurance premium given the predicted probability of relatively infrequent (but economically costly) events. We do not know if the data required for these calculations exist. The comparison between states of the world (with different maps of landuse, including the replacement of areas of extensive with intensive agriculture) would allow for an estimate of which breeds/varieties would be more likely to be lost given their spatial distributions. Particularly relevant would be to understand which of these breeds/varieties occur in specific areas and are not safe somewhere else (e.g. in seed banks). The economic value of those losses would be calculated from the estimate of the ‘premium’ values of each breed/variety. We suspect that this model would be best for calculating changes in crop, rather than livestock, genetic diversity, given that the latter are less likely to be directly related to changes in landuse. For breeds/varieties conserved ex-situ (e.g. in seed banks) the relevant comparison would be between one scenario in which the creation and maintenance of these banks is ensured versus one where it is not — the costs of creation/maintenance would then be compared with the benefits (predicted avoided costs due to disease and pest outbreaks). 69 Adequacy of scenarios We predict that the key data that scenarios would have to produce is information on landuse change, including on the intensification of agriculture. Scenarios could also specifically assess the costs and benefits of maintaining breeds and varieties ex-situ. 4.3.7 Insights for economic valuation The general model proposed would already provide results in economic terms. 4.3.8 Some key resources Food and Agriculture Organization of the United Nations (FAO), Commission on Genetic Resources for Food and Agriculture. Has published both The State of the World’s Plant Genetic Resources for Food and Agriculture (1997) and The State of the World's Animal genetic Resources for Food and Agriculture (2007). International Treaty on Plant Genetic Resources for Food and Agriculture (ITPGRFA) Global crop diversity trust Convention on Biological Diversity (CBD) — Agricultural Biodiversity International Food Policy Research Institute (IFPRI) International Livestock Research Institute (ILRI), that has build the database Domestic Animal Genetic Resources Information System _(_DAGRIS), containing research-based information on the distribution, characteristics and status of 871 breeds of cattle, sheep, goats, pigs and chickens indigenous to Africa and Asia. Consultative Group on International Agricultural Research (CGIAR) is a strategic alliance of countries, international and regional organizations, and private foundations supporting 15 independent international agricultural research centers. Bioversity is one of these CGIAR-supported Centres. Bioversity International (formerly IPGRI) is one of those centes. Germplasm Resources Information Network (GRIN) within the U.S. Department of Agriculture's Agricultural Research Service. Possible experts to contact include: 70 e Melinda Smale, Senior Research Fellow at IFPRI. Her research emphasizes the development of methods to assess the value of crop biodiversity and the identification of policies to enhance the utilization and management of crop genetic resource, particularly in developing economies. e Jan Engels, Genetic Resources Science and Technology Group, at Bioversity International. e José Esquinas-Alcazar, FAO Commission on Genetic Resources for Food and Agriculture. 4.3.9 Participants Authors Hannah Peck, Andrew Balmford and Ana Rodrigues (University of Cambridge, UK) 71 4.4 Soil quality for crop production This section is a very brief overview based on a quick literature review. It greatly benefited from expert contributions (see below) but it has not been reviewed by experts in this field. There seem to be four main ways in which wild nature contributes to the benefits that we obtain from agricultural crops (e.g. food, fibres, fuels) by improving or retaining farmland soil quality. We called these: internal effects (soil biota); conversion effects (when non- agricultural land is converted to agriculture); neighbourhood effects (when neighbouring, non-agricultural, systems contribute to soil quality in croplands); and wild fertilisers (e.g. guano). Note that the importance of soil for water regulation is covered separately (section 4.9). Internal effects The soil fauna and flora, and the communities occupying the agricultural land, affect the soil properties and therefore the quality of the soil for agriculture. These ‘internal effects’ seem to be the main way in which wild nature affects soil quality, by affecting: nutrient fixation (e.g. nitrogen fixation by bacteria); nutrient cycling (e.g. in organic matter decomposition, by fungi, bacteria, dung beetles...); soil structure (e.g. by plant roots, termites...); water regulation, particularly water holding capacity and drainage (e.g., by plant roots, termites, micro-organisms and earthworms...); uptake of water and nutrients (mycorrhizal fungi); erosion regulation (e.g. by vegetation cover and leaf litter); suppression of pests and diseases (e.g. by mycorrhizal fungi). More than species diversity, it seems that functional diversity (and its influence in trophic interactions) is key to the decomposition, nutrient cycling, and stability of soil processes. Soil diversity is important for resilience to stress and disturbance (Griffiths et al. 2000; Brussaard et al. 2007). As agricultural intensification occurs, the regulation of functions through soil biodiversity is progressively replaced by regulation through chemical and mechanical inputs (Giller et al. 1997; Swift et al. 2004). The type of agricultural management is the primary determinant of the composition, diversity and functioning of the soil food web. So, to reap the rewards of soil biodiversity in farming systems requires identifying those practices that best harness the beneficial roles of the soil food web. One idea commonly proposed is that management regimes that encourage a soil community that bears the closest resemblance to natural ecosystems are most likely to require fewer inputs because of greater reliance on ecosystem self-regulation (e.g. Bardgett & McAlister 1999; Mader et al. 2002). In tropical agricultural systems undergoing intensification, large numbers of farmers have limited access to inputs, and therefore the maintenance and enhancement of soil biodiversity may be particularly relevant to such farmers (Giller et al. 1997). One key feature of natural ecosystems is a soil community that is dominated by fungal-based pathways of decomposition, whereas intensive agriculture tends to promote more bacterial-based pathways of decomposition. Therefore, farming systems, for example non-tillage systems, cessation of fertiliser and organic farming, that encourage fungi and their consumers might reap additional benefits in terms of nutrient cycling, enhancement of plant diversity through the action of mycorrhizal fungi (Smith et al. 72 2008), carbon storage (e.g. Chivenge et al. 2007), and nutrient retention under stress (e.g. Gordon et al. 2008). We predict that the importance of ‘internal effects’ is increasing as agricultural land expands and appropriate soil management becomes more important to ensure long-term crop production (particularly when the inputs of chemical fertilizers and pesticides are restricted for environmental reasons), and as organic agriculture expands (Mader et al. 2002). We do not know the extent to which the economic consequences of different agricultural management regimes on crop yields (specifically through their effect on soil quality) have been quantified. We get the impression that while a fair amount of work seems to have been done in comparing organic and conventional farming systems, this has mainly focused on temperate regions. Even if information is available on how crop yields are affected by changes in the soil biota resulting from different management regimes, it would still be difficult to assess these effects globally. Indeed, the scenarios being produced to compare a business-as- usual world with a more biodiversity-friendly world would need detailed spatial information not only on the specific crops but also on the management regime for each crop. We suspect this would not be possible for Phase 2. Other relevant references: e Cassman et al. (2005) — section on cultivated systems of the Millennium Ecosystem Assessment (volume Ecosystems and Human Well-being: Current States and Trends). e Bloem et al. (2004) — Measuring soil biodiversity: experiences, impediments and research needs. e Bloem et al. (1997) — Food webs and nutrient cycling in agro-ecosystems e Bloem et al. (2006) — Microbiological methods for assessing soil quality e Schjonning et al. (2004) — edited volume on Managing Soil Quality: Challenges in Modern Agriculture, particularly section by Brussaard on Biological soil quality from biomass to biodiversity — Importance and resilience to management stress and disturbance. e FAO (1999) — Sustaining Agricultural Biodiversity and Agro-ecosystem Functions. e Brussaard et al. (2007) — Soil biodiversity for agricultural sustainability e Wood et al. (2000) — Pilot analysis of global ecosystems: Agroecosystems Conversion effects These are one-off inputs from wild nature, when non-agricultural land is converted into agricultural land. This may happen as agriculture expands into areas occupied by natural ecosystems (e.g. the Amazonian frontier), or as a cyclic events through shifting agriculture. Here, crops benefit from the improvements in soil quality that were accumulated over time in the previous form of land use (e.g. long-term soil improvement in forest systems). 73 Agricultural expansion into new areas often occupies terrains that are nor particularly suitable for agriculture, and soil fertility may decline very quickly as crops effectively mine the soil nutrients (Alfsen et al. 2001; Carr et al. 2006). Soil fertility may however be retained under appropriate agricultural practices that ensure adequate replacement of the nutrients used by the crops (Barber & Diaz 1994). Shifting cultivation, also called ‘swidden’ agriculture or ‘slash-and-burn’ agriculture, is one of the oldest forms of farming and consists of cropping on cleared plots of land, alternated with lengthy fallow periods. These systems are the dominant form of agriculture in tropical humid and subhumid upland regions and are typically associated with tropical rain forests. Shifting cultivation is practiced on about 22% of all agricultural land in the tropics and is the primary source of food and income for some 40 million people (Giller and Palm 2004), including many of the world’s poorest. While the contribution to global food security is negligible, given the low yields and general lack of infrastructure in areas where shifting cultivation predominates, this method of cultivation has a potentially large impact on regional and global ecosystem services through its effects on biodiversity, greenhouse gas emissions, and soil nutrients. Although these systems are generally associated with soils of low fertility, they are highly sustainable and resource-conserving in areas with low population density (Kleinman et al. 1995). However, high population density increases the pressure on available land and resources, reducing the time available for a regenerative fallow between cropping cycles (Cassman et al. 2005), and effectively converting the region into an agricultural frontier. We predict that the importance of conversion effects is increasing as the agricultural frontier expands in regions such as the Amazon. On the other hand, it is expected to decline as population increases, rendering shifting agriculture non-viable. In comparing different states of the world (created by different scenarios) it should be possible to account for the economic value of soil fertility for crops in agricultural frontiers, quantified as one-off subsidies in the conversion from natural ecosystems to agriculture. As for shifting agriculture, by knowing the time needed for soil recovery and the number of people that can be sustained per unit of area of cultivated plots it should be possible to map potential for sustainable shifting cultivation worldwide and the maximum population densities allowed for long-term sustainability. This would then be crossed with data on demand (regions where populations are more likely to depend on this type of agriculture) as a basis to calculate the potential economic value of a sustainable inputs of natural ecosystems to farmland soil through shifting agriculture. Given current human population densities, it is predicted that shifting cultivation can only provide a minor contribution to food production (from a global perspective), although as mentioned poor people in some regions will depend on this type of agricultural practice. The time needed for field to recover is likely to be variable and will depend on soil type and climate. Other relevant references: e Moreira et al. (2006) — book Soil Biodiversity in Amazonian and Other Brazilian Ecosystems. ¢ Noordwijk (1995) — book Below-ground Interactions in Tropical Agroecosystems. 74 Neighbourhood effects When neighbouring, non-agricultural, systems contribute to improve or maintain soil quality in croplands. We envisage that this could include, for example, coastal sand dunes sheltering inland agricultural fields from sea salt spray, or uphill forests protecting downslope croplands from erosion (by reducing water runoff) or from landsides. The importance of neighbourhood effects may be increasing as agriculture expands from the most favourable lowlands into slopes, or declining as the size of fields expands and therefore reduces the contact with other ecosystems. Assuming that these effects are essentially related to the physical protection of farmland soil by neighbouring ecosystems, it seems conceptually possible to model this effect, such that, for example, one would be able to quantify the changes in farmland soil quality in a watershed having either more or less coverage of upslope forest. However, we found no evidence for the existence of such models. Wild fertilisers These are products obtained from wild nature, which are used as crop fertilisers, for example: guano, seaweed, peat and fish. These subsidies to soil quality have been very important historically (e.g. guano was the basis of the Peruvian economy in the mid 19" Century) but their importance has declined substantially with the development of artificial fertilisers. Wild fertilisers may still be important in areas where farmers cannot afford artificial fertilisers, though. We suspect there is not a centralised dataset on current use of wild fertilisers that could be used as the basis of for a spatially explicit economic evaluation. 75 4.4.1 Can we quantify and map the contribution of wild nature for the provision of soil quality for crops, and how it might change? What can be done in Phase 2? By whom? Internal effects We predict that it is unlikely that even a first-cut quantification could be possible of the value of wild nature for agriculture through internal effects. Indeed, the type of agricultural management is the primary factor in this evaluation, and so modelling the economic consequences of different policy options on soil fertility (and crop yields) would require detailed information not only on how specific crops vary across space but also on management practices in each plot. We suspect this is not a feasible task for Phase 2. Conversion effects We predict that it should be possible to obtain a first-cut model for quantifying the value of wild nature for agriculture through conversion effects, both frontier and shifting cultivation. We suggest further expert consultation (see below) to evaluate the feasibility of this particular task and understand who the appropriate experts would be. Neighbouring effects While conceptually possible to quantify, we found no evidence for researchers thinking along these lines and therefore suspect that this task would not be feasible in the short-term. Wildlife fertilisers We suspect there is not a centralised dataset on current use of wild fertilisers that could be used as the basis of for a spatially explicit economic evaluation on Phase 2. 4.4.2 Some key resources FAO Land and Water Development Division, particularly its section on soil health for food security. They maintain the Soil Biodiversity Portal: The role of soil organisms in agriculture for better land management decision. ISRIC — World Soil Information, including David Dent (Director) and Zhanguo Bai (global assessment of land degradation and improvement). Alterra, Wageningen University, The Netherlands, including: Peter de Ruiter, Jaap Bloem and K.B. Zwart Kenneth G. Cassman and John W. Doran at the University of Nebraska-Lincoln, USA. Stanley Wood at the International Food Policy Research Institute, Washington DC, USA. Rattan Lal at the School of Natural Resources, College of Food, Agricultural, and Environmental Science, Ohio State University, USA. 76 e Peter Kleinman at the Department of Crop and Soil Sciences, Pennsylvania State University, USA e Paul Mader at the Research Institute of Organic Agriculture, Switzerland e Richard D. Bardgett Soil and Ecosystem Ecology, Lancaster University, United Kingdom e Sara Scherr at Ecoagriculture Partners, USA 4.4.3 Participants Authors Ana Rodrigues, Antares Hernandez, Andrew Balmford (University of Cambridge, UK) Contributors The following experts provided very valuable insights on which this section was based. We apologise to them that time constraints prevented us from circulating the text for revision. Kenneth G. Cassman (University of Nebraska-Lincoln, USA) Jaap Bloem (Alterra, Wageningen University, The Netherlands) Richard D. Bardgett (Lancaster University, UK) Acknowledgements Stanley Wood (IFPRI, Washington DC, USA) provided useful references. 77 4.5 Livestock This section is a very brief overview based on a quick literature review. It has not been reviewed by experts in this field. 4.5.1 Why is the production of livestock important for human wellbeing? Livestock and livestock products (e.g. milk, eggs) are estimated to make up over half of the total value of agricultural gross output in industrial countries, and about a third of the total in developing countries (Wood et al. 2005). Globally, it accounts for 40% of agriculture gross domestic product, provides one-third of humanity protein intake, employs 1.3 billion people and creates livelihoods for one billion of the world’s poor (Steinfeld et al. 2006). Livestock are also a significant source of income and consumption in developing countries, often providing a supplementary source of income and income stability, and frequently one of the few assets available to the poorest. Livestock production also affects human wellbeing negatively by: causing land degradation and desertification, habitat loss, climate change (18% of greenhouse gas emissions), and water pollution (see Steinfeld et al. 2006 for a very comprehensive review). 4.5.2 What are the overall trends in the production of livestock? The global importance of livestock and their products is increasing with population growth, changes in food preferences (associated with rising incomes and urbanisation), and expanding global trade. Global production of meat and meat are projected to more than double between 2000 and 2005 (Steinfeld et al. 2006). 4.5.3 How is the production of livestock affected by changes in wild nature? Livestock production is the single anthropogenic user of land either directly through grazing (26% of the ice-free terrestrial land surface) or indirectly through consumption of fodder and feedgrains (33% of total arable land). In all, livestock production accounts for 70% of all agricultural land and 30% off all land surface area of the planet (Steinfeld et al. 2006). An increasing fraction of livestock production is intensive, with crops provided as feed to animals kept at high densities. Globally, it is estimated that 35% of all cereal use are fed to animals. In this case, the contribution of wild nature for livestock production is indirect, via the contribution of nature to crop production, including through pollination (section 4.1), biological control of crops (section 4.2), soil quality for crops (section 4.4) as well as through water regulation (section 4.9) and the regulation of natural hazards (section 4.13). An important fraction of livestock production takes place in natural grasslands, and that is the focus of this section. Here livestock production is being directly subsidised by the primary productivity of natural ecosystems, and the species and ecosystem processes that sustain it. Wild nature also subsidises livestock production when pasturelands expand into other ecosystems. Expansion of livestock production is a main driver of deforestation, particularly in Latin America — 70% of the previously forest landed area in the Amazon is 78 occupied by pastures, and feedcrops occupy much of the rest (Steinfeld et al. 2006). Here, wild nature is contributing by providing soil fertility, as discussed for the conversion effects in section 4.4. The contribution of genetic diversification to livestock production is discussed in section 4.3. Relationship between habitat area and livestock production We expect this relationship to be linear. 4.5.4 What are the main threats to livestock production in natural grasslands? Overexploitation of natural grasslands, by keeping livestock at a higher density than what is sustainable, leads to degradation of the grasslands that is not easily reversed. For example, about 20% of the world’s pastures and rangelands, with 73% of rangelands in dry areas, have been degraded to some extent, mostly through overgrazing, compactation, and soil erosion created by livestock action. Drylands tend to be particularly affected as they have lower natural productivity and livestock is often the only source of livelihoods to people living in those areas (Steinfeld et al. 2006). Climate change is likely to affect livestock production, both by affecting natural rangelands (particularly in regions where predictions are of increased aridity) and by affecting feed crop production. 4.5.5 Are abrupt changes likely in the provision of livestock in natural grasslands? We found no evidence for future abrupt and disproportionate changes in the capacity of natural systems (grasslands) to produce livestock. 4.5.6 Can we quantify and map the contribution of wild nature for livestock production, and how it might change? A first-cut map of the contribution of wild nature for livestock production already exists, as a map of the annual production of livestock derived — at least in part — from grazing on unimproved natural grasslands (Naidoo et al. in press). This was obtained by combining global data on livestock distributions (Wint & Robinson 2006), producer prices, and current and potential vegetation (Ramankutty & Foley 1999; Bartholome & Belward 2005) To map livestock production on natural pastures, Naidoo et al. (in press) used recently- developed 5° resolution global maps of livestock distributions (Wint & Robinson 2006), which used regression-based methods to estimate the expected density of cattle, sheep, goats, pigs, poultry, and buffalo across the earth’s surface. For each livestock type, they used these density estimates and data on the mass of edible meat per animal (estimated by country from FAOSTAT data) to estimate the tons of meat produced in each cell. A global producer price was used to weight different livestock types; using these weights, an aggregate index of livestock production was obtained by summing the weighted livestock meat weights. Naidoo et al. (in press) then constructed a global map of natural pastures by combining a 5’ resolution potential vegetation data set of savanna, grassland/steppe, and shrubland biome types (Ramankutty & Foley 1999) and then masking out all known human-altered landscapes using the Globa! Land Cover 2000 remote sensing data 79 (Bartholome & Belward 2005). They intersected the livestock production index and the map of natural pastures to produce a global map of livestock production on natural pastures. Limitations of these data include the difficulties of mapping pastures from remotely sensed imagery and the lack of spatially explicit weightings that reflect differences in the economic values of livestock species in different regions of the world (Naidoo et al. in press). The new pastureland maps by Ramankutty et al. (2008) provide more updated information than Ramankutty & Foley (1999). The map by Naidoo et al. (in press) provides an estimate of current production in grasslands. However, this may overestimate the natural sustainable production rates. On the one hand, existing livestock may in some areas receive other food supplements, and therefore not depend completely on the natural grassland production. On the other hand, livestock may be kept at densities that are too high, therefore degrading the natural grassland productivity in the long term (Steinfeld et al. 2006). Hence the Naidoo et al. (in press) map is not ideal, but we believe it is a very useful first cut to quantifying the value of wild nature for livestock production in natural grazing areas. In the longer-term, it would be useful to calibrate this map to account for the maximum natural productivity of land. Adequacy of scenarios Appropriate scenarios would need to predict landuse and demand for livestock products. 4.5.7 Insights for economic valuation The map produced by Naidoo et al. (in press) has as units an aggregate index of livestock production, but it could have equally have been produced in terms of estimated monetary value of the livestock produced in each cell. 4.5.8 Some key resources - Wint & Robinson (2006) — Gridded Livestock of the World, by the FAO. - Ramankutty et al. (2008) present newly derived global maps of agricultural lands, including pasturelands. - Steinfeld et_al. (2006) present an assessment of the full impact of the livestock sector on development and the environment, produced by the Livestock, Environment and Development (LEAD) Initiative of the FAO. 80 4.5.9 Participants Authors Ana Rodrigues, Andrew Balmford (University of Cambridge, UK) 81 4.6 Marine fisheries This section is a fully developed review, including contributions by experts in the field, some of whom subsequently reviewed the text. 4.6.1 Why are marine fisheries important for human wellbeing? This section focuses on marine capture fisheries, not covering marine aquaculture. Marine capture fisheries are a globally important source of food: fish is highly nutritious, rich in micronutrients, minerals, essential fatty acids and proteins, and represents a valuable supplement to diets otherwise lacking essential vitamins and minerals. Globally fish provides more than 2.8 billion people with at least 20 percent of their average per capita animal protein intake (FAO 2006). Marine capture fisheries contribute 56% of the total amount of fish available for human consumption (FAO 2006). Marine fisheries also affect the provisioning of food indirectly, via the use of fishmeal and fish oil in aquaculture and in poultry and pig production (Arthurton et al. 2007). Demand for fish is increasing with population growth, rising wealth and changing food preferences as a result of the marketing of fish in developed countries as part of a healthy diet (Arthurton et al. 2007). Accordingly, marine capture fisheries are an important source of economic benefits, with an estimated first-sale value of US$84.9 billion (FAO 2006), and important for income generation, with an estimated 38 million people employed directly by fishing, and many more in the processing stages (Arthurton et al. 2007). Marine fisheries are increasingly valuable for recreation, particularly in developed countries. In the US alone, in 2006 nearly 13 million anglers made more than 89 million marine recreational fishing trips on the Atlantic, Gulf and Pacific coasts, capturing almost 476 million fish, of which 55 percent were released alive (Van Voorhees & Pritchard 2007). In the European Union (EU 15), an estimated 8 million recreational sea anglers (Stevenson 2007) spend an estimated €25 billion a year, compared to a €20 billion value for commercial landings in 1998 (Pawson et al. 2004). 4.6.2 What are the overall trends in the provision of marine fisheries? The global trends in marine fisheries have been reviewed in the Millennium Ecosystem Assessment (Pauly et al. 2005). Catches increased substantially during the twentieth century as fishing fleets worldwide expanded. These trends continued until the 1980s, when global marine landings reached slightly over 80 million tons per year; then they either stagnated or began to slowly decline. However, regional landings peaked at different times throughout the world, which in part masked the decline of many fisheries. This global decline is despite an increase in fleet size, effort, and technology, in conjunction with an expansion into previously unfished areas, depths and stocks (Pauly et al. 2005; Morato et al. 2006). It is estimated that in 2005 around one-quarter of the stock groups monitored by FAO were underexploited or moderately exploited (3% and 20%, respectively) and could perhaps produce more. About half of the stocks (52%) were fully exploited and therefore producing catches that were at or close to their maximum sustainable limits, with no room for further expansion. The other one-quarter were either overexploited, depleted or recovering from 82 depletion (17%, 7% and 1%, respectively) and thus were yielding less than their maximum potential owing to excess fishing pressure exerted in the past, with no possibilities in the short or medium term of further expansion and with an increased risk of further declines and need for rebuilding (FAO 2006). The real extent of collapses in marine systems is likely to be more dramatic than realised, as massive declines that took place before systematic statistics began are likely to go unnoticed as perceptions are adjusted to what a natural marine system is (the ‘shifting baseline syndrome’; Carlton et al. 1999; Jackson et al. 2001b). A meta-analysis of collapsed fisheries has indicated minimal recovery even after fishing moratoria are imposed (Hutchings 2000, Hutchings and Reynolds 2004). Loss of production has mean loss of earnings which has been disportionately absorbed by developing countries whose resources were the last to be fished down by multinational fleets (Srinivasan et al. 2008). 4.6.3 How is the provisioning of marine fisheries affected by changes in wild nature? Biomass of the exploited species is the key parameter determining catches in marine fisheries, at least in the short-term. Global declines in exploited biomass are inferred from declines in catches despite increased effort. In the North Atlantic, for example, current overall biomass of high-trophic level fishes is estimated to be one third of what it was in 1950 (Christensen et al. 2003). Fishing pressure has reduced the biomass of some species to less than 10% of the pre-exploitation level within a few decades, particularly species with vulnerable life history traits such as large predatory fishes (Myers and Worm, 2003), including sharks and relatives (Myers et al. 2007), and deep sea species (Devine et al. 2006). While commercially exploited species have been particularly affected (e.g. Cod, Cook et al. 1997; bluefin tuna, Fromentin & Powers 2005), non-target species have also suffered depletions as by-catch (e.g. leatherback turtles, Spotila et al. 2000). Fishing efforts tend to be concentrated on high-trophic species first. As these become depleted, the biomass of lower-trophic species may increase due to predator release (Scheffer et al. 2005). Fishing efforts then move down the food web, and landings from global fisheries have accordingly shifted in the last few decades from large piscivorous fishes toward smaller invertebrates and planktivorous fishes, especially in the Northern Hemisphere (Pauly et al. 1998). This may at first lead to increased catches, but overfishing of lower-trophic species will also result in their declines, and so fishing down the food web is unsustainable in the long-term (Pauly et al. 2005). Provisioning of fishmeal for aquaculture and for meat production can be done using lower trophic levels species but recreational fisheries benefit specifically from large, long-lived, high-trophic fish species. There is evidence that species diversity is important for marine fisheries, both in the short- term, by increasing productivity, and in the long-term, by increasing resilience. A recent review of fisheries trends in large marine ecosystems (Worm et al. 2006) found that the proportion of collapsed fisheries decayed exponentially with increasing species richness. Furthermore, the average catches of non-collapsed fisheries were higher in species-rich systems. Diversity also seemed to increase robustness to overexploitation, with rates of recovery positively correlated with fish diversity. This increased stability and productivity was attributed to a portfolio effect (Tilman et al. 2006), whereby a more diverse array of species provides a larger number of ecological functions and economic opportunities, leading to a more stable trajectory and better performance over time (Worm et al. 2006). 83 The diversity of genetically distinct populations, adapted to particular regions, is also likely to be important, especially in terms of the sustainability and resilience of wild stocks to longer-term natural and anthropogenic change (Hilborn et al. 2003a; Worm et al. 2006; Hiddink et al. in press). Maintaining the genetic diversity of marine fisheries keeps also options open for future farming of marine species. Given that marine fisheries are sustained by lower-trophic species, they depend directly on the biomass of other, non-targeted, species, including phytoplankton and zooplankton. The diversity of non-target species, such as prey species, may also be important: experiments have shown the importance of diverse food sources for secondary production and for the stability of populations across trophic levels in marine ecosystems (Worm et al. 2006). The removal (or depletion below a certain level) of populations of particular species or functional groups has been shown to have dramatic effects on some marine ecosystems and the associated fisheries. Predators in particular (‘top-down control’) seem to be very influential in shaping and maintaining various habitat states or population levels (Myers et al. 2007). Many examples of regime shifts associated with biodiversity losses have been documented (reviews in: Jackson et al. 2001b; Bellwood et al. 2004; Folke et al. 2004; Agardy et al. 2005) and include: kelp forest losses due to urchin population explosions after declines in large predatory fishes and sea otters; die-offs of turtlegrass beds following turtle population collapse; sudden catastrophic coral mortality due to overgrowth in algae after grazers (herbivorous fishes, urchins) have been lost (Mumby et al. 2007); trophic cascades following the collapse of cod and other large-bodied predators in the northwest Atlantic, transforming a highly productive system dominated by benthic fish into systems dominated by pelagic fish and macroinvertebrates with poor fisheries productivity (Frank et al. 2005); and two successive regime shifts in the Black Sea triggered by overfishing and compounded by euthrophication, firstly after depletion of top predators and secondly after depletion of planktivorous fish, resulting in a jellyfish-dominated, fisheries-poor system (Daskalov et al. 2007). Recovery following change may be slow and may follow a different trajectory from the one observed during the decline (hysteresis; Hughes et al. 2005). Fisheries records provide compelling evidence of dramatic declines in the stocks of many exploited species. Substantially less is known about trends in non-targeted species. Marine species have long been assumed to be resilient to extinction, and therefore until recently been neglected in extinction risk assessments (Baillie et al. 2004). This is now changing, with ongoing assessments demonstrating that as human activities take over the world’s oceans (Halpern et al. 2008) marine systems are suffering serious declines in species diversity, both locally (particularly in coastal habitats; Jackson et al. 2001b; Worm et al. 2006) and globally (e.g. predatory fish communities; Myers and Worm 2003). The largest ongoing assessment of species conservation status is the Global Marine Assessment (due to assess 20,000 marine species by 2010), which is assessing the threat status of every marine vertebrate species, and of selected habitat-forming invertebrates and plants. In addition, experimental evidence suggests that a loss of biodiversity increases vulnerability to the establishment of invasive species (Stachowicz et al 1999, 2007; Worm et al. 2006). Marine fisheries are vulnerable to the decline in the extent or quality of particular marine habitats that play important roles in the provisioning of key resources (e.g. food, shelter) for targeted species. These include, amongst others: 84 Coral reefs, which directly support fisheries that constitute 9-12% of the world’s total fisheries (up to 25% in some parts of the Indo-Pacific), providing livelihoods for millions of people in tropical coastal regions. A large number of offshore fisheries also rely on the supporting services of reefs as breeding, nursery or feeding grounds (Moberg and Folke 1999; Agardi et al. 2005). Cold water reefs are also important to fisheries (Fossa et al. 2002; Freiwald et al. 2004), although much less is known about these. Coral reefs are one of the world’s most threatened habitats, particularly in the Caribbean (Gardner et al. 2003). Coral reefs are suffering a worldwide decline in quality and extent (Pandolfi et al. 2003; Bellwood et al. 2004, Newton et al. 2007), with some regions particularly degraded (e.g. Caribbean, Gardner et al. 2003; SE Asia, Middle East and Indian Ocean, Wilkinson 2000). They are also one of the habitats most at risk from climate change (Hoegh-Guldberg et al. 2008). Seamounts are very poorly known systems, but thought to be important for the provision of food, as sites for spawning aggregations, and nurseries/refuges for juvenile open ocean fish (Moore & Jennings 2000; Stocks 2004). Out of an estimated 100,000 seamounts (>1 km high), only about 350 have been sampled biologically (many more have no doubt been fished out), and less than 100 of these have been studied in any detail; the ones that were revealed extraordinary levels of species endemism (EuroCoML 2008). Seamounts are among ‘newly’ targeted ecosystems that have been intensively fished since the second half of the 20th century, particularly through highly destructive bottom trawling (Watson & Morato 2004). However, their long-lived, slow- growing, species are particularly vulnerable to over-exploitation, and so reports of drastic collapses are accumulating (Devine et al. 2006). Whole populations have been known to be depleted in just a few years (Devine et al. 2006) while few deep-sea fisheries have recovered from bottom trawling even two or three decades after fishing ceased (Morato et al. 2004; Watson & Morato 2004). Seagrass meadows are highly productive systems, thought to have a fundamental role in maintaining populations of commercially exploited fish and invertebrate species, in both tropical and temperate regions, by providing a permanent habitat, a temporary nursery area for the development of the juvenile stages, a feeding area for various life- history stages, and/or a refuge from predation (McArthur et al. 2003; Agardy et al. 2005). They are also thought to maintain fisheries indirectly by providing organic matter that is incorporated into coastal nutrient cycles (Jackson et al. 2001a; Gullstrom et al. 2002). Major losses of seagrass habitat have been reported from the Mediterranean, Florida Bay, and Australia (Duarte 2002). Present losses are expected to accelerate, especially in Southeast Asia and the Caribbean (Burke et al. 2001; Duarte 2002), as eutrophication increases, algal grazers are overfished, and coastal development increases. Kelp forests are highly productive temperate systems, with a complex biological structure organized around large brown algae, supporting a high diversity of species and species interactions. Kelp support fisheries of a variety of invertebrate and finfish species, particularly by providing nursery and shelter habitats, and the kelp itself is harvested for food and additives. Overfishing of predator species has led to regime shifts in many regions towards systems that are totally dominated by urchins and of substantially less value for fisheries (Agardy et al. 2005). Marine fisheries are also vulnerable to the declines in the extent or quality of coastal habitats, including: 85 Mangroves are found in intertidal zones and estuarine margins in tropical and subtropical regions, where they can be important to fisheries as refuge areas from predators or physical disturbance, or via food input (Agardy et al. 2005). Mangroves are particularly important for supporting artisanal fisheries in some regions; in Bangladesh, for example, the artisanal fishery (contributing 85-95% of the total coastal and marine catch) is highly influenced by mangroves (Islam & Haque 2004). Furthermore, mangroves may be ecologically linked to seagrass beds and coral reefs, with synergistic effects for fish productivity (Mumby et al. 2004). Much of the coastal population of the tropics and sub-tropics resides near mangroves, and it has been estimated that 35% of mangrove forests have been loss in the recent past principally due to conversion to agriculture, aquaculture, salt pans, and urban and industrial development (Valiela et al. 2001; Agardy et al. 2005). Estuaries and coastal wetlands play a key role in marine fisheries provision, particularly as nursery areas for fisheries (Beck et al. 2001), but also by providing habitat for commercially important molluscs and crustaceans, and in regulating water quality (Agardy et al. 2005). The loss of these habitats has been linked to the collapse of fisheries in North America, North Africa and elsewhere, although further quantitative studies are required (Agardy et al. 2005). Estuaries and coastal wetlands have been degraded, altered, or eliminated in many areas, particularly through reduction of the estuaries watershed through agricultural conversion, eutrophication, pollution by pesticides and herbicides, overfishing, and biological invasions (Agardy et al. 2005). Marine fisheries are furthermore affected by changes in inland ecosystems that affect the quality, volume and timing of water inputs as well as erosion regimes: Forests, wetlands and riparian habitats play an important part removing excess nutrients from freshwater runoffs (see section 4.9). This role is becoming increasingly important as agricultural intensification is increasing nutrient loads, leading in some regions to eutrophication responsible for extensive “dead-zones’ that can impact seriously on the production of fisheries (Rabalais et al. 2002; Agardy et al. 2005). The volume and timing of freshwater inputs to marine systems affect the salinity in estuarine and other coastal environments, in turn affecting the fisheries that depend on them. Inland changes in the regulation of the timing and volume of freshwater input are therefore likely to affect the breeding and recruitment of fisheries species dependent these environments. A dramatic example of fisheries declines associated with changes in freshwater regime can be found at the mouth of the Colorado River (Kowalewski et al. 2000; Rowell et al. 2005). Coral reefs, seagrass meadows, mangroves and estuaries can be degraded or even destroyed by increasing sedimentation caused by landuse change, and so inland erosion regulation is important for the conservation of these habitats and of the species that rely on them (Agardy et al. 2005). For example, a decline in the tropical fisheries of the Caribbean and Pacific has been linked to the degradation of coral reefs, seagrass beds, and mangroves from sedimentation (Rogers 1990). Relationship between habitat area and fisheries provisioning This question can be addressed in at least three ways: i) Are larger areas of suitable habitat (e.g. larger coral reefs, larger mangroves) more productive in terms of fisheries per unit of area than smaller ones? 86 ii) How does fisheries productivity in a region vary with the fraction of area reserved? iii) Are larger reserves more effective at improving fisheries than smaller ones? Answering the first question is important to understanding whether the effects on fisheries yields are proportional to area change when habitat declines (e.g., when part of a coral reef is dredged). We found only one study that directly investigated what a fisheries-area relationship looks like (Watson et al. 1993) but there are likely to be more. We suggest that the relationship is likely to be linear at least for larger areas, such that doubling habitat area doubles overall yields (Figure la). However, we also predict that there is likely to be an area threshold (Roberts et al. 2003), below which yields decline disproportionately with habitat loss. Whether the threshold corresponds to a large (purple line) or a small area (blue line) is likely to depend on the extent to which there is low or high species redundancy for fisheries production, and the extent to which the system has top-down regulation. For high redundancy and bottom-up regulation, fisheries per unit area are likely to depend essentially on the primary productivity of the habitat, and will likely remain unaffected by habitat loss up to very small areas, when area finally becomes too small to retain viable populations of any exploited species. For systems with low redundancy and/or strong top-down community regulation, a threshold effect is likely to take place as the area tips below the minimum viable area needed to support the most sensitive targeted species and/or keystone species. As discussed above, there is evidence that marine systems are at least in some cases affected by strong top-down regulation (e.g., by predators, large herbivores; Jackson et al. 2001b; Folke et al. 2004). There is also evidence that increasing species diversity increases fisheries productivity (Worm et al. 2006), indicating low functional redundancy (Micheli & Halpern 2005). This suggests that there may be a relatively high threshold in the relationship between area and fisheries provisioning (Figure la: purple line). These thresholds are likely to become even higher when considering the long-term resilience of fisheries productivity. We found no empirical tests investigating how the fraction of a region that is reserved (in no-take zones) affects overall fishing productivity in the region. Theoretical metapopulation models (Botsford et al. 2003) suggest that the effect will vary with the overall fishing effort. For relatively low fishing effort, high catches are obtained even in the absence of reserves; they can be improved by a low level of reservation, but beyond that an increase in reserve area reduces catches as it reduces the area available for fishing. For high fishing efforts, though, low levels of reservation may be unable to avert population collapses, and therefore null catches (this is because the movement of animals and cheating of fishers around boundaries depletes stocks near the boundaries even within a reserve, and so if the reserve is too small there is no protection rendered); in this case, it may pay to set aside a substantial fraction of the region as reserves. We did not find studies that investigated directly how the size of individual marine reserves (no-take zones) affects fishing yields in the surrounding areas. There have however been several studies that measured differences in fish communities (diversity, biomass and/or density) inside and outside protected areas. It is assumed that positive differences result in higher spill-over effects resulting in improved fisheries in the surrounding areas (Roberts et al. 2001; Halpern 2003). Recent reviews have investigated whether these differences vary with reserve size: most found no relationship (Coté et al. 2001; Halpern 2003; Guidetti & Sala 2007) but one found a positive relationship (Claudet et al. in press). If there is no relationship with area, then many small reserves are predicted to be a better option for maximising fisheries within a region (as they provide higher edge:area ratios and therefore 87 more gain from spills over for the same area), although larger reserves are likely to be needed to ensure long-term viability (Roberts et al. 2003). Even if in general there is no relationship between reserve effect and area (Figure 1b, blue line), there is likely a threshold below which the effect declines very quickly with area (see discussion above) below which the effect declines very quickly with area, for the same reasons as described above. The size of such threshold has not been determined (Roberts et al. 2003). Reserve effect may also plausibly increase with reserve size, as overall species richness increases (Neigel 2003), but it is likely that it does so with diminishing returns (Figure 1b, purple line). a) b) D ‘= oO So oO sd wn 5 BER x} S55 oO C+ oo pas Ow Qa ®o®co 7p) Og 5 @ cme = o-~ oO © oH oO s £ OD ' -Es 1 {e\ : > ? habitat area reserve area Figure 1. Hypothetical relationship between habitat area and fisheries production, represented in three ways. a) Relationship between area of a given habitat and fisheries production, for lower (blue) and higher (purple) thresholds. b) Relationship between fraction of a region reserved and fisheries production across the entire region, for low (blue) and high (purple) fishing effort. c) Relationship between the size an individual reserve and reserve effect (difference between fish biomass/density inside and outside the reserve, assumed to be an indicator of fisheries catches), assuming no relationship (blue) but an area threshold, or a positive relationship (purple). 4.6.4 What are the main threats to the provision of marine fisheries? Human activities have now affected all oceans to various degrees (Halpern et al. 2008). Marine fisheries are under threat from different, but often interdependent and self- reinforcing, pressures (Pandolfi et al. 2005). Overfishing is the most widespread and the dominant direct impact on marine fisheries (Pauly et al. 2005). Fishing pressure is increasing as technological advances allow industrial fleets to fish with greater efficiency, farther offshore, and in deeper waters to meet the global demand for fish. Overfishing is in many parts of the world stimulated by subsidies which either reduce the costs of fishing or increase the net revenue fishers obtain (Pauly et al. 2005). Illegal, unregulated, and unreported ([UU) fishing is estimated to constitute a substantial fraction of the overall catches, undermining efforts for sustainable fisheries management (Pitcher et al. 2002). 88 Habitat destruction affects the provisioning of fisheries by marine systems. Most is caused by highly destructive fishing techniques, including bottom trawling and coral reef dynamiting. Expansion of coastal infrastructure may affect marine habitats directly (e.g., dredging during port construction) or indirectly (e.g. by increasing sedimentation). Underwater mining activities (minerals, gas and oil) pose an increasing threat. Some systems may recover fairly quickly, but others, such as cold-water corals and seamounts, may take hundreds of years to recover (Pauly et al. 2005). Pollution is particularly prevalent in coastal habitats, including through sewage discharge and agricultural runoff. In some regions, the extremely high nutrient loads trigger hypoxia conditions known as dead zones (Malakoff 1998; Rabalais et al. 2002). The open seas can also be affected by pollution, for example oil spills. Past waste dumping, particularly radioactive waste, poses a significant threat to deep-sea ecosystems should the containers leak, which seems likely over the long term (Pauly et al. 2005). There is still poor understanding of the effects of persistent organic and inorganic pollutants on marine fisheries (Pauly et al. 2005). Climate change is predicted to have both direct and indirect impacts on marine fisheries. Direct effects act on physiology and behaviour and alter growth, development, reproductive capacity, mortality, and distribution of targeted species. Indirect effects alter the productivity, structure, and composition of the ecosystems on which fish depend for food and shelter. Some of the changes are expected to have positive consequences for fish production, but in other cases negative impacts are predicted (Brander 2007). Ocean acidification is predicted to have dramatic impacts for calcifying organisms, including coral reefs (Hoegh-Guldberg et al. 2007) as well as some types of zooplankton, particularly at high latitudes (Orr et al. 2005). Other current threats include biological invasions, such as by competitors (e.g. Daskalov et al. 2007) or pathogens (e.g. Mumby et al. 2007), while future threats may include larger- scale projects for carbon sequestration through iron enrichment (Pauly et al. 2005). 4.6.5. Are abrupt changes likely in the provision of marine fisheries? Fisheries yields of individual species are well-known to be subject to sudden collapse following overexploitation, with stocks of slow-growing, slow-maturing species in particular often failing to recover to former levels of abundance (e.g., Hutchings 2000; Watson & Morato 2004). There are many documented examples of recent sudden regime shifts in marine systems (e.g. Jackson et al. 2001b; Bellwood et al 2004; Frank et al. 2005; Daskalov et al. 2007), with implications for fisheries provisioning. Such shifts seem to be particularly likely in ecosystems that are or have been under intense fishing effort, and which have been simplified by the loss of one or more higher-trophic functional groups. Many apparently stable systems have very little resilience, with shifting between stages possibly depending on small changes in the abundances of one or a very few species (e.g., Mumby et al. 2007). While the collapse of entire fisheries has been observed across relatively large areas (e.g. Black Sea; Daskalov et al. 2007) more often the collapse of a particular species or set of species results in a shift in fishing effort towards other species (often further down in the food web; Pauly et al. 1998) or towards other regions/ecosystems (e.g. towards increasing depths; Pauly et al. 2005). These shifts mask the underlying sequential collapses from 89 ocean-level or global fisheries statistics. Under current knowledge, it is therefore unlikely that a synchronised global collapse will be observed by 2025, but it is very likely that the steady decline that has been observed since the mid-1980s will continue, with some local- or regional-sclae collapses (Pauly et al. 2005). Climate change (Brander 2007) and related ocean acidification (Orr et al. 2005) are the greatest sources of uncertainty in predictions of marine fisheries, and potentially could be responsible for sudden, large-scale, changes in the foreseeable future. 4.6.6 Can we quantify and map the global provision of marine fisheries and how it might change? State of knowledge and data availability Fisheries management has traditionally relied on single-species models aimed at guiding harvesting decision for particular species of interest (Hilborn et al. 2003b). In the last two decades, there has been a shift towards ecosystem-based fisheries management, stimulated by the development of ecosystem modelling (Christensen & Walters 2005). These are particularly relevant as tools for evaluating scenarios and trade-offs associated with particular policy options, and as such particularly relevant for the Review on the Economics of Biodiversity Loss. A diversity of models have been developed and these are reviewed and compared in Plaganyi (2007). One of the most widely used models is the Ecopath with Ecosim (EWE) modeling approach (Christensen & Walters 2005), with development spearheaded by the University of British Columbia’s Fishery Centre (www.ecopath.org). EWE combines software for ecosystem trophic mass balance analysis (Ecopath), with a dynamic modeling capability (Ecosim) for exploring past and future impacts of fishing and environmental disturbances as well as for exploring optimal fishing policies. Recent versions of the software have brought Ecosim much closer to traditional single-species stock assessment, by allowing age-structured representation of particular, important populations and by allowing users to ‘fit’ the model to data. Ecosim models can be replicated over a spatial map grid (Ecospace) to allow exploration of policies such as marine protected areas, while accounting for spatial dispersal/advection effects and migration (Christensen & Walters 2004). EwE has already been applied for regional-scale policy exploration aiming to achieve economic, social and ecological sustainability objectives (Pitcher & Cochrane 2002). 90 Based on the EwE approach and software, the University of British Columbia’s Fishery Centre has just developed a global marine model to explore scenarios for the world’s oceans (Alder et al. in press). EcoOcean is spatially defined by the 19 FAO fishing areas covering the world’s oceans and driven by effort of five fleets. The model was constructed using 43 functional groups, selected with special consideration for exploited fish species but intended to include all major groups in the oceans. The model relies on global fisheries datasets available through the Sea Around Us Project (SAUP) and FAO websites, in particular fish catches, ex-vessel prices, and fleet statistics. The model output from EcoOcean can be used to describe how landings, profits, the marine trophic index and depletions may change under different policy scenarios in different areas of the world. For the first time, EcoOcean provides a common reporting platform so that the outcomes of the different scenarios can be compared within and between geographic areas, as well as for fleets and fisheries. Fishing effort is the most important driver for the ecosystem model simulations, and so EcoOcean can be applied to investigate the consequences for fisheries of policy actions under different scenarios, such as a variation in the area covered by marine reserves, or the implementation or not of fishing quotas for particular groups of species. It also tracks changes in community structure, for example in the Mean Trophic Index (Pauly et al. 1998) However, currently the finest spatial resolution possible for the scenario comparison is that of the 19 FAO fishing areas used to develop the model. The model is therefore not appropriate for investigating the consequences of finer-scale changes, such as the implementation of particular marine reserves, the loss of a coral reef, or local changes in water quality as a result of terrestrial runoff, for example leading to the creation of dead zones. The next version of EcoOcean has a planned spatial resolution of half-degree cells (Alder et al. in press), which will allow greater sensitivity to these changes. Further ongoing developments in the quality of the dataset by the SAUP will improve the accuracy and reduce the uncertainty associated with the models. These limitations notwithstanding, EcoOcean seems a good candidate for evaluating the consequences of policy measures towards reducing biodiversity loss on the provisioning of marine fisheries. Indeed, EcoOcean has already been applied to explore the scenarios posed in the Global Environment Outlook 4 (UNEP 2007) and the International Assessment for Agricultural Science, Technology and Development, demonstrating its usefulness as a policy tool. The outputs from these two assessments have provided policy makers with plausible results on which to base future decisions regarding management of fisheries and marine ecosystems. For other models of marine ecosystem management, see Plaganyi (2007). Adequacy of scenarios Adequate scenarios for application of marine ecosystem management models must specify fishing effort across space, and for each functional group. 91 4.6.7 Insights for economic valuation Appropriate scenarios will generate different states of the world, with corresponding predictions of fisheries intensity across the world’s oceans, for different functional groups. EcoOcean can be used to investigate how these translate into long-term trends in fisheries yields for each state of the world. Data on the market value of those fisheries, and appropriate discount rates, can be used to monetise the difference in fisheries yields between different states of the world. This market value should reflect predicted variations in demand for different regions/functional groups. For example, increasing demand in fishmeal for aquaculture may result in increasing prices for small, pelagic fish. An increase in recreational fisheries, on the other hand, may lead to increasing value in the stocks of large, high-trophic species. 4.6.8 Some key resources e Fishery Centre at the University of British Columbia, Vancouver, Canada, is a leading centre of research on global marine systems, including the development of modelling tools such as Ecopath with Ecosim (EwE), a freely available ecological/ecosystem modelling software suite. The Fishery Centre is also the home of the Sea Around Us Project (SAUP) that actively collates and analyses data on global fisheries catches, fishing effort and costs of fishing to investigate the impact of fisheries on the world's marine ecosystems to support the development of sustainable, ecosystem-based fisheries policies. SAUP is led by Daniel Pauly and includes amongst other leading experts Jackie Alder, Villy Christensen, and Reg Watson. e The Food and Agriculture Organisation (FAO) Fisheries and Aquaculture Department collates and distributes global data on catches, value, consumption, fleets and trade for marine fisheries. e United Nations Environment Programme (UNEP) Global Marine Assessment: following from a call at the World Summit on Sustainable Development (WSSD) for the "establishment by 2004 and under United Nations of a regular process for global reporting and assessment of the state of marine environment, including socio- economic aspects, both current and foreseeable, building on existing regional assessments", and subsequent approval by the United Nations General Assembly, > This is our recommendation for Phase II, based on the results of this review. It does not commit the leaders of Phase II to follow it, and it does not commit the recommended research group to actually do such work. 92 effort is underway to evaluate all the recent assessments at regional and national scales for fisheries as well as other marine resources/issues. A first report Survey of Global_and Regional Marine Environmental Assessments and Related Scientific Activities has been completed by UNEP-WCMC. e A Global Environmental Facility funded Large Marine Ecosystems Strategy is underway for the assessment and management of international coastal waters, as a joint effort between the IUCN -— The World Conservation Union, the Intergovernmental Oceanographic Commission of UNESCO (IOC), other United Nations agencies, and the US National Oceanic and Atmospheric Administration (NOAA). This programme is led by Ken Sherman at NOAA. e The Global Marine Species Assessment is based at Old Dominion University, VA, USA, coordinated by Kent Carpenter. A partnership between the IUCN Species Survival Commission and Conservation International’s Centre for Applied Biodiversity Science, this project will be the first global review of the conservation status of every marine vertebrate species, and of selected invertebrates and plants. The project involves a range of partners in compiling and analyzing all existing data on approximately 20,000 marine species, and will determine the risk of extinction according to the IUCN Red List Categories and Criteria. The goal is to complete Red List assessments of approximately 15,000 marine species by the year 2010. e The Census of Marine Life a global network of researchers in more than 80 nations engaged in a 10-year scientific initiative to assess and explain the diversity, distribution, and abundance of life in the oceans. The world's first comprehensive Census of Marine Life-past, present, and future-will be released in 2010. e FAO Fisheries and Aquaculture Department provides statistics on fisheries commodities and trade, global production, fishing fleets, fishers, and consumption of fish and fisheries products. e SAUP is actively collating and making freely available data on global fisheries catches, fishing effort and costs of fishing. e The Stock Recruitment Database consisting of maps, plots, and numerical data relating to over 600 fish populations (over 100 species) from all over the world, collated by the late Ransom Myers at the Department of Biology at Dalhousie University in Halifax, Nova Scotia, Canada. e FishBase is was developed at the WorldFish Center in collaboration with the FAO and other partners. FishBase is a relational database with information on the ecology, distribution, and conservation of ~28,500 marine and freshwater species (practically all fish species known to science). 4.6.9 Participants Authors Kelly Flower, Ana Rodrigues, Andrew Balmford (University of Cambridge, UK) Reg Watson (SAUP, University of British Columbia, Canada) 93 Contributors/Authors (currently reviewing) Jackie Alder (SAUP, University of British Columbia, Canada) Boris Worm (Dalhousie University, Canada) Acknowledgements Daniel Pauly and Rashid Sumaila (SAUP, University of British Columbia, Canada) provided contacts and suggestions for additional experts. 94 4.7 Inland fisheries and aquaculture This section is a developed overview based on a literature review and expert consultation. It greatly benefited from expert contributions (see below) but it has not been reviewed by experts in this field. 4.7.1 Why are inland fisheries and aquaculture important for human wellbeing? Inland capture fisheries and aquaculture provide globally important sources of food and income. Catches from inland waters, about 90 percent of which occur in Africa and Asia, account for 27% of global fisheries production, including include 9.6 million tonnes (7%) from capture fisheries and 28.9 million tonnes (20%) from aquaculture (70% of which come from China; FAO 2006). They employ more than 50 million people directly and many more indirectly (Finlayson et al. 2005). These fisheries are almost entirely finfish, with some regional exploitation of crustaceans and molluscs (Wood et al. 2005). Three types of freshwater fisheries are covered in this section: Subsistence and artisanal fisheries are important for their roles in nutrition, food security, income generation and informal employment. They are particularly important in certain regions, such as SE Asia and Africa, where they can be the only source of animal protein (Finlayson et al. 2005). These fisheries rely on a broad diversity of species and corresponding diversity of fishing gear (Coates et al. 2003). In Asia, many people practise integrated agriculture-aquaculture, maintaining and fishing the fish fauna of rice paddies as an additional income or food source (Dugan et al. 2006). There are very few reliable official statistics available, and the available figures are thought to be large underestimates (Finlayson et al. 2005; FAO 2006). For example, Coates et al. (2003) estimates the total fisheries production of the Lower Mekong Basin to be between 2.6 and 21 times higher than official statistics. This is mainly due to the diffuse nature of small-scale fisheries (both temporally and spatially), with data often based on household surveys and creel census; and the fact that many rivers and lakes are part of more than one country (FAO 2006). Many African countries have poor statistical information, despite the importance of inland fisheries to their economies and food supply — for example, in 2004 the FAO had to estimate the catch for over half of the countries where inland fisheries were known to occur (FAO 2006). Commercial fisheries are mainly centred on the Great Lakes of North America and Africa, although some do occur on the world’s major rivers (Finlayson et al. 2005; Coates et al. 2003). For example, the Great Lakes of North America have supported one of the largest freshwater fisheries for a century or more, valued (along with sport fishing) at $4 billion annually (Finlayson et al. 2005). Freshwater commercial fisheries are mostly locally traded, with a few species (e.g. salmon) making it to the international market. In addition, developed countries tend to trade within themselves, whereas developing countries export mostly to developed ones (FAO 2006). Within commercial fisheries, there is a marked trend towards aquaculture-based systems, as inland water systems (particularly lentic) are managed more intensively via feeding and restocking activities to produce a higher yield (‘agriculturalisation’ of freshwater systems; Wood et al. 2005). Often, there is a focus on a few productive species, rather than the ecosystem as a whole. For example, the Great Lakes fisheries in North America consist of a mix of native and introduced species, which are maintained by regular stocking of cage- reared individuals (Finlayson et al. 2005), whereas the Lake Victoria fishery, which was once based on over 500 species, now relies on the exotic Nile Perch for ~90% of their catch (Njiru et al. 2008). There is a continuum of increasing levels of food and stocking input in freshwater systems — from purely capture fisheries to high-input, high-technology aquaculture — that makes these two types of fisheries difficult to separate in practice. Recreational fisheries are dominant in northern temperate countries, where they contribute significantly to the economy. In Europe, for example, the inland recreational fishing industry has been valued at $25 billion/year. This move towards recreational fisheries is also increasing in developing economies (e.g. Brazil, India, Argentina, and several states of the Zambezi River basin), often to boost the local tourism economy (Dugan 2006). These fisheries tend to focus on a few prize species, with a substantial proportion being catch- release. In some areas, there is an increasing trend to manage freshwater systems (including through stocking, feeding, and habitat management) to favour recreational species, particularly in reservoirs (Finlayson et al. 2005). 4.7.2 What are the overall trends in the provision of inland fisheries and aquaculture? Catches from inland waters, have shown a slowly but steadily increasing trend (about 2% a year) since 1950, owing in part to stock enhancement practices. However, there are different trends across regions. Catches are declining in Europe, Canada and the United States; for example, Europe has seen a 30% decline in catches since 1999 (FAO 2006). In some cases, increases are due to better recording of subsistence fisheries, for example in the Lower Mekong Basin (Dugan et al. 2006). Aquaculture continues to grow more rapidly than all other animal food-producing sectors, with an average annual growth rate for the world of 8.8 percent per year since 1970. However, there are signs that the rate of growth for global aquaculture may have peaked, although high growth rates may continue for some regions and species (FAO 2006). Recreational fisheries are increasing as a result of urbanisation, increasing affluence and larger populations, in both developed and developing economies. A significant proportion of the catch from inland waters in developed countries is from recreational fisheries (FAO 2006), and they too have the potential for dramatic impacts on inland fish stocks (Cooke & Cowx 2006). It is thought that subsistence and artisanal fisheries are in decline, due to a pressure to commercialise and the shrinking of rural populations. However, trends are difficult to assess due to the paucity of reliable official statistics. 4.7.3 How is the provision of inland fisheries and aquaculture affected by changes in wild nature? The provisioning of inland fisheries ultimately depends on the biomass of targeted species. In most systems this is sustained ‘bottom-up’ by the productivity of primary producers, which in turn are largely influenced by physicochemical factors (however, some deep lake demersal fisheries also depend on detrital fallout). ‘Top-down’ trophic interactions at higher levels also regulate primary productivity and maintain conditions for higher trophic level species (Carpenter 1987; Carpenter & Kitchell 1988). 96 The overall increases in inland fisheries catches mask the underlying decline in the biomass of many individual stocks as most inland fisheries in the developing world are overexploited (Finlayson et al. 2005; Wood et al. 2005; Dugan et al. 2006). As with marine fisheries, there has been a decline in the mean trophic level of inland fish catches over the last few decades (‘fishing down the food web’; Welcomme 1999; Wood et al. 2005). This tends to result in the replacement of larger, slower-reproducing species towards smaller, fast-reproducing species (including invertebrates), whilst the total productivity may remain constant or even increase initially as predators disappear (Welcomme 1999). Hence individual species declines are masked as fisheries switch to other species, as well as by the use of enhancements (such as stocking or feeding) which are increasingly required to maintain yield, and by the general rising production of all types of aquaculture (Finlayson et al. 2005; FAO 2006). Hence, there is no evidence for a generalised collapse in inland fisheries, even if there have been instances of local collapses related to modifications of freshwater systems, such as dam building (e.g. the loss of migratory guilds after the building of a dam on the Mekong; Amornsakchai et al. 2000). There does not seem to be much evidence for a relationship between species diversity and freshwater fisheries yields. Indeed, increases in productivity have been observed in some systems despite dramatic declines in diversity; for example, catches from Lake Victoria increased from 30,000 mt to an average of 500,000 mt since the 1970s, despite the loss or near-loss of about half the native species, with the introduced Nile Perch now making around 90% of the landing’s volume and value (Njiru et al. 2008). However, it is likely that species diversity (both of targeted species and others) improves the resilience of capture fisheries and aquaculture to anthropogenic and environmental changes, as seems to be the case with marine fisheries (Worm et al. 2006; section 4.6). Indeed, inland fisheries are being simplified to depend on a very few species that do well in the modified habitats that we are creating, and while this seems to be working well for fisheries in the short-term it may be risky in the long-term if conditions (e.g. climate) change. In the transition to more intensively managed freshwater fisheries, the tendency is to focus on a few more desirable and/or more productive species. However, the genetic diversity of managed populations remains very important, both as a source of variation for breeding in desirable traits and to increase the resilience of farm fish to disease. Also, there is still a very broad set of freshwater species being domesticated/selected for aquaculture production, and so there is a greater reliance on species diversity than in other food production sectors (FAO 2006). A diversity of species is also important for the colonisation of new habitats (such as reservoirs) created by freshwater modification. Often, species from the previous habitats (e.g. flowing water environments) are unable to establish themselves, and in this case other species are frequently introduced, such as cichlids and carps (Finlayson et al. 2005). Particular keystone species can be very important in some freshwater systems. For example, the loss of andromadous salmon from rivers results in a substantial reduction in nutrient input (Allan et al. 2005; also McIntyre et al. 2007) and beavers’ activities substantially change freshwater habitats (Rosell et al. 2005; Collen & Gibson 2001). 97 Freshwater biodiversity is highly threatened, including for example 56% of endemic fish in the Mediterranean basin (Smith & Darwall 2006), and 54% of endemic fish in Madagascar (Darwall et al. 2005). Freshwater communities are highly affected by loss, degradation, and transformation of freshwater habitats (Finlayson et al. 2005). Furthermore, many freshwater species use distinct habitats for breeding, hunting and nurseries (e.g. hundreds of species migrate from large rivers/lakes to streams in order to breed). This renders them particularly vulnerable, as they can be affected by the degradation of any of those habitats, as well as by factors affecting the linkages between habitats, such as through the construction of dams and through the separation between the main river channel and the floodplain. Communities in lotic systems (rivers, streams) are frequently dependent on particular hydrological regimes, which can be disturbed by land use changes (e.g. logging), water abstraction, and dam building (Finlayson et al. 2005; Welcomme & Halls 2004). These systems are also affected by river canalisation, separation from the flood plain, and water pollution. Lentic systems (lakes, reservoirs, wetlands) are particularly vulnerable to eutrophication (e.g., in Lake Victoria a doubling of nutrient input led to an 6-8 times increase in algal biomass; Balirwa 2007), sedimentation (e.g. in Lake Malawi forest loss resulted in increased sedimentation, and reduced quality habitat for cichlid species; Rusuwa et al. 2006), the introduction of invasive species (e.g. Nile perch in Larke Victoria; Njiru et al. 2008; Balirwa 2007), water abstraction (e.g., Aral Sea; Finlayson et al. 2005), and loss/decline of area due to land use change (e.g. Lake Victoria has lost many wetland habitats at a current rate of ~3.8% annually; Njiru et al. 2008). Floodplains are key habitats for many species, particularly for breeding, and they are being threatened by disconnection from main river channel and by habitat loss as the floodplain is appropriated for agriculture and human settlements. Changes in the habitats within the water catchment area of freshwater systems (including riparian habitats) can have important and widespread effects on the quality, extent and composition of freshwater ecosystems (Finlayson et al. 2005). Terrestrial systems can contribute to the regulation of energy and material transfer between the riparian and aquatic ecosystems, particularly in small water bodies (Pusey & Arthington 2003), affecting water quality (e.g., via leaf litter input; Allan & Flecker 1993), water turbidity and sedimentation (e.g. increasing erosion following deforestation and road building; Alin et al. 1999), and hydrologic regimes (e.g., and changes in quantity and seasonality of water flow following deforestation; Allan and Flecker 1993; Welcomme & Halls 2004). Terrestrial habitats can also provide nutrition to freshwater species (e.g. fruit eating fish guilds) and contribute to the structure of freshwater habitats (e.g. by affecting sedimentation and the presence of wood debris; Wright & Flecker 2004). While the negative effects of habitat changes on freshwater communities are well- established, it is often less obvious what the effects are on freshwater fisheries proper. In some cases, the creation of new habitats favours particular species, which in turn can be exploited in fisheries. Examples include reservoirs following dam construction, and rice fields following the conversion of floodplain habitat (Dugan et al. in press). In many cases, it would seem that some species benefit whilst others lose out, and so the advantages/disadvantages for fisheries are dependent on the role of species in the ecosystem, and their importance to fisheries. 98 Relationship between habitat area and the provision of inland fisheries We predict that in principle marine fisheries decline linearly with the area of relevant habitat (e.g. floodplain, rivers). However, reality is likely to be much more complex, given the intrinsic interconnectivity of freshwater systems. Hence, a stretch of river cannot be retained on its own if the upper stream river is lost, while a floodplain area cannot exist in isolation from the main river channel. 4.7.4 What are the main threats to the provision of inland fisheries and aquaculture? Water abstraction (particularly for agriculture) is perhaps the main threat to inland fisheries and aquaculture, altering flowing regimes and reducing river flow in riverine systems (sometimes completely, as in sections of the Colorado River), reducing the area of available habitat in lentic systems (e.g. Aral Sea, where water volume has been reduced by 75% since 1960; Finlayson et al. 2005), and allowing saline water intrusion in coastal freshwater systems (Finlayson et al. 2005). Water translocation is also an increasingly important consideration, as water is diverted from one river basin to another (e.g. Yellow River in China). Fisheries are directly impacted as the flood cycle and water availability are disrupted, and there is an increased threat from invasive alien species as river basins are linked. River_canalisation and separation from the floodplain changes hydrologic regimes substantially, affecting communities in both river systems and floodplains. Floodplains are also directly threatened by habitat loss, as the land is appropriated for agriculture and human settlements. Indeed, floodplain systems are one of the most threatened habitats in the world; Europe, for example, has lost up to 95% of its riverine floodplains, and those that are left are often decoupled from the hydromorphological dynamics of the river (Tockner & Stanford 2002). Dam building destroys habitats upstream, causes major disruption to hydrological and sedimentation flows downstream, creates a barrier to species migrations, and fragments resident species populations. Fisheries based on migratory species may be totally compromised (e.g. sturgeon fisheries). However, fisheries may improve in the reservoir area when there is restocking with suitable (frequently alien) species (Finlayson et al. 2005). Overexploitation leads to population reductions and even extinction of the targeted species. A FAO major assessment of inland fisheries (1999, in Finlayson et al. 2005) reported that most inland capture fisheries that rely on natural reproduction of the stocks are overfished or are being fished at their biological limit. Water pollution and eutrophication (particularly due to agriculture, but also urban and industrial run-off) significantly affect freshwater communities. Eutrophication, caused by increased nutrient loads (particularly due to fertilisers in agriculture runoff) is the most widespread problem in lakes and reservoirs, and also one of the most difficult to abate. Toxic contaminants (e.g. heavy metals and organochroline compounds) may render fisheries inappropriate for human consumption (Finlayson et al. 2005) Landuse changes in terrestrial habitats (e.g. deforestation) affect water and sediment regime flows and water quality. 99 Invasive species may affect fisheries directly, by predating, competing with, or parasiting the targeted species, or indirectly, by affecting habitat characteristics (e.g. water hyacinth) and water regimes (e.g. mimosa; Finlayson et al. 2005) Climate change is likely to affect inland fisheries and aquaculture by altering hydrologic flows. Many aquaculture practises also contribute to the degradation of inland water systems, via pollution, the spread _of diseases and the introduction of alien species (Finlayson et al. 2005). It is important to keep in mind that some anthropogenic changes are also resulting in increased fisheries, due to stocking and feeding (particularly of small lakes and reservoirs) and management of habitat structure for particular target species (Finlayson et al. 2005; Lake Victoria example: Njiru 2008). 4.7.5 Are abrupt changes likely in the provisioning of freshwater fisheries and aquaculture? As discussed above, there does not seem to be evidence of widespread collapse in freshwater fisheries as a result of overharvesting. It is not impossible that the ‘fishing down the food web’ process may lead to such collapses, as is happening in marine systems (section 4.6), but on the other hand. freshwater systems may follow a different trajectory given the tendency towards increased management (feeding, restocking etc). Human-induced eutrophication can trigger sudden shifts in lakes and reservoirs from clear to turbid due to algal blooms. These blooms may include toxic cyanobacteria and affect freshwater fisheries and recreational use of the water bodies. Reduction of nutrient concentrations is often insufficient to restore the original state, with restoration requiring substantially lower nutrient levels than those at which the regime shift occurred (Scheffer et al. 2001). Particular species invasions may also cause localised disruption to fisheries and aquaculture. Climate change has perhaps the potential to cause widespread disruption to inland fisheries and aquaculture, but currently this is essentially speculation. 4.7.6 Can we quantify and map the global provision of inland fisheries and aquaculture, and how it might change? There is currently no overall global model for the productivity of freshwater fisheries and aquaculture. Different models exist for different systems in different areas, including for example: 100 e Tropical river floodplain — model investigating the impact of hydrological changes on fish yield/productivity, combining a hydrological and population model to simulate production changes with different water regimes (Halls & Welcomme 2004). e Northern temperate lakes — the Morphoedaphic Index (Ryder 1965) has been extensively used as a predictor of fisheries yields based on total dissolved solids and depth of lake. However its accuracy is contested, and there are many subtle variations in analysis. e African water bodies — models which predict fisheries production have been generated by the Africa Water Resources Database. These models use physico- chemistry and climatological maps to predict production based (mainly) on surface area. It is likely that different freshwater systems will need to be modelled separately. As with marine fisheries (section 4.6), it is key that models reflect sustainable production, rather than current catches, as the latter are likely to be unsustainable in the long term. Except for some well-studied systems, the state of knowledge regarding freshwater fisheries is very poor. FAO has historically managed global statistics, but they are currently decentralising for freshwater fisheries and so they are not the central repository anymore. The leading role in freshwater statistics is now taken by World Fish Centre. The best available data are being compiled at regional scales (e.g. Mekong River Commission Fisheries Programme). 101 What can be done in Phase 2? At what cost? By whom?? The experts we consulted were of the opinion that a reasonable model, or at least a first estimate, could be generated within one year, by reviewing existing studies on sustainable freshwater fisheries production across the world and extrapolating from those. Two possible tiers were proposed: a) Engage a group of 12-15 experts covering different regions, including people with good skills in GIS and statistical modelling (see suggested names under key resources), based on existing models and statistics, ask them to provide a predictive collective opinion of how different states of the world would differ in inland fisheries and aquaculture production. One or several collective meetings (NCEAS-style workshops) would be needed. As in a), but proceeded by regional meetings to collect and collate existing (but currently dispersed) regional information on which to base finer estimates for productivity change. It is estimated that it would take around 36-48 researcher-months for a suitable team to compile a reasonable global estimate from existing knowledge (tier a). Resources would need to be allocated for meetings. Adequacy of scenarios In order to be useful in predicting the impacts of changes in wild nature on the provision of freshwater fisheries and aquaculture, the scenarios would ideally provide information on: e The availability of water (temporally and spatially), particularly affected by abstraction for agriculture and hydroelectric projects. e Land available in the floodplain, closely linked to the expansion of agriculture and urban development. e The connectivity of water systems, particularly the construction of dams. e The canalisation of rivers. 4.7.7 Insights for economic valuation The subsequent economic evaluation, based on the maps of biophysical production, would need to carefully consider what the end use of each type of fisheries is. Different valuation methods would be necessary in each case: 3 This is our recommendation for Phase II, based on the results of this review. It does not commit the leaders of Phase 2 to follow it, and it does not commit the recommended research group to actually do such work. 102 For subsistence and artisanal fisheries, data are likely to come from household surveys and creel census (i.e. Asia, South America, Africa) For commercial fisheries, most data can be taken from the market in fish and fish produce (African and North American Great Lakes, some large rivers). Recreational fisheries are likely to need valuation using data from licenses and expenses, utilising Contingent Valuation and Travel Cost methods (North America, Europe, Australia, some parts of the tropics). Some key resources Patrick Dugan, at the World Fish Centre, would be in a good position to oversee this project. R. Hilborn, at the University of Washington (USA) and E. Jeppesen (Denmark) could be valuable with regards to statistical modelling of fisheries-ecosystem relationships. Robin Welcomme, Institute of Fisheries Management, UK. Will Darwall, Freshwater Biodiversity Unit, SSC/IUCN, Cambridge, UK: leading the global freshwater assessment. Peter Mcintyre, School of Natural Resources and Environment, University of Michigan, USA. Gerd Marmulla, FAO. Robin Abell, Senior Freshwater Conservation Biologist, Conservation Science, WWE-US, Washington DC, USA. Carmen Revenga, World Resources Institute, Washington DC, USA. Regional expertise: o Mekong/Asia: Eric Baran (CGIAR), Chris Barlow (Mekong River Commission Fisheries Programme); Ashley Halls (Aquae Sulis Ltd, UK). o Africa: Patrick Dugan, Africa Water Resources Database. o Latin America/South America: Miguel Petrere (Universidade Estadual Paulista, Rio Claro, SP, Brazil; for Brazil and much of Latin America); Oriana Almeida (IPAM, Amazonian Institute of Environmental Research, Brazil; for the Amazon); Angelo Antonio Agostinho (Universidade Estadual de Maringa, PR, Brazil). o Europe: tan Cowx (University of Hull, UK) o Australia: Angela Arthington (Griffith University, Australia) 103 4.7.9 Participants Authors Kelly Flower, Andrew Balmford, Ana Rodrigues (University of Cambridge, UK) Contributors The following experts provided very valuable insights on which this section was based. We apologise to them that time constraints prevented us from circulating the text for revision. Will Darwall (SSC/IUCN, Cambridge, UK) Patrick Dugan (World Fish Centre) Peter McIntyre (University of Michigan, USA) Robin Welcomme (Institute of Fisheries Management, UK) Acknowledgements Robin Abell (WWF-US) provided useful suggestions for experts. 104 4.8 Wild animal products This section is a fully developed review, including contributions by experts in the field, most of whom subsequently reviewed the text. 4.8.1 Why is this benefit important for human wellbeing? The harvesting of wild animal products is defined here to include the capture of terrestrial wild animals or their products, including mammals, birds, reptiles, amphibians and invertebrates (for example snails and insects, including honey harvesting). Wild animal harvesting can be broadly classified into three categories: subsistence, commercial, and recreational (Table 2). In subsistence harvesting, the benefits obtained from wildlife (particularly food) are directly consumed or used by, and play a very significant role in the subsistence of, the harvester and its family (e.g. Peres 2000). Commercial harvesting takes place when most of the products are sold for profit (e.g. caiman meat trade; Silveira et al. 1999; chameleon pet trade; Carpenter et al. 2004). Recreational hunting refers to activities in which the main objective is the personal enjoyment of the hunter, rather than food or profit (e.g. trophy lion hunting in Tanzania; Whitman et al. 2004). There is substantial overlap between these categories. The differentiation between subsistence and commercial harvesting may be subtle (with, for example, subsistence hunters often selling the excess or the most valuable species as a source of income; Wilkie & Carpenter 1999b), and the transition between the two may happen quickly as markets open (e.g. Sierra et al. 1999). Recreational hunting may also have roots in traditional (either subsistence or commercial) hunting activities (McCorquodale 1997). Table 2 Comparison of main attributes of subsistence, commercial and recreational hunting. — Attribute Subsistence hunting Commercial hunting Recreational hunting Purpose Food Profit Sport Users Rural people, poor Rural and urban High income Volume High Very High Low Area Often close to residence, Distant from residence Distant from residence, permanent variable Frequency Year round Year round Weekends, hunting seasons Principle Prey Mostly mammals Mostly mammals Mostly birds, but medium to large mammals also. Legality Variable according to Often illegal Usually legal country Control Difficult but may be Often very difficult Feasible feasible Social Value _ High Low Medium The main benefit obtained from the harvesting of wild animals is arguably food (e.g. Fa et al. 2003), with other benefits including raw materials such as hides (e.g. Iriarte & Jaksic 1986), medicines or substances traditionally considered to have medicinal value (e.g. Alves & Rosa 2005; Ntiamoa-Baidu 1997), pets (e.g. Carpenter et al. 2004), and personal 105 enjoyment (e.g. Wilkie & Carpenter 1999a). Here we focus particularly on harvesting for food (referred throughout as ‘wild meat’) in tropical regions, with some discussion of harvesting for personal enjoyment (‘recreational hunting’). Wild meat is a valuable source of nutrition across the globe, but its importance as food is particularly high in tropical regions (with some forest-living peoples obtaining up to 90% of animal protein from wild animals; Fa et al. 2003). The literature focuses on large vertebrates (> | kg), which seem to contribute the most to wild meat provision (particularly mammals, followed by birds, reptiles and amphibians) and this is the emphasis of this section as well. Invertebrates also have a very significant nutritional role in some areas, but these are frequently ignored in studies of wild food harvesting, and traditional consumption is declining due to a negative bias by Westerners (deFoliart 1999). deFoliart (1999) presents a comprehensive review of the use of insects as food across the world, including for example a study that found that 65 species of insects furnished 10% of the animal protein intake in the Democratic Republic of Congo (compared to 30% from game and 47% from fishing), and the case of Thailand, where more than 80 species of insects (35 families) are used as food, and locust and grasshopper catching are becoming alternatives to pesticide spraying during plague periods. Wild meat is important for the poor and landless, particularly during times of the year when agricultural production is lower (de Merode et al. 2004), and periods of famine and insecurity or conflict, when normal food supply mechanisms are disrupted and local or displaced populations have limited access to other forms of nutrition (Elliott et al. 2002; Wood et al. 2005). In some regions, though, reliance on wild meat is permanent (see Ntiamoa-Baidu 1997 for a review of wild meat use across Africa). While the dependency on wild meat is greatest in rural communities, townspeople can also be major wildlife consumers (Milner-Gulland et al. 2003). In the transition from rural to urban consumption, wild meat consumption becomes less of a nutritional need (in which case people will readily switch to domestic meat, if it becomes more readily available) and more a matter of preference based on tradition or status (in which case at least some consumption persists even when wild meat becomes substantially more expensive than domestic alternatives) (Milner-Gulland et al. 2003; see Goudarzi [2006] for an extreme example of African wild meat being sold as a luxury product in US and European markets). Urban demand creates the opportunity for trade, and therefore wild meat harvesting is also an important source of income for local people (e.g. Silveira et al. 1999; Ntiamoa-Baidu 1997). Estimates of the annual wild meat harvest, mainly mammals, include 23,500 tonnes in Sarawak, 67,000— 164,000 tonnes in the Brazilian Amazon, and 1—3.4 million tonnes in Central Africa (Milner-Gulland et al. 2003 and references therein). Wild meat harvest and trade are often excluded from official statistics (Wood et al. 2005) but the economic value of annual trade has been estimated at, for example, over US$175m for the Amazon Basin and US$200m for Céte d’Ivoire (Rao & McGowan 2002 and references therein). Recreational hunting is the dominant form of terrestrial wild animal harvesting in the most economically affluent countries. The nutritional value of wild meat is negligible given wide access to cheaper alternatives. And, quite the opposite of being a source of income, hunting becomes a substantial expense as hunters spend considerable sums for the pleasure and social status of doing it. Regions in Europe and the United States with higher fractions of rural population tend to have a higher frequency of hunters, supporting the hypothesis that the motivation to hunt is based on cultural roots (Heberlein et al. 2002). In the United States, 5% of the population 16 years old and older went hunting in 2006, spending US$22 106 billion (USFWS 2007). In the European Union, there are more than seven million hunters (FACE 2007), estimated to spend €10 billion per year (Pinet 1995). Trophy hunting by foreign tourists generates at least US$201 million per year in sub-Saharan Africa, from a minimum of 18,500 clients (Lindsey et al. 2007), while in Eurasia US$33-39 million dollars are generated from 45,000 to 60,000 foreign hunters (Hofer, 2002). Harvesting of wild animals has costs to human wellbeing as well. It is widely acknowledged that overexploitation is resulting in a worldwide depletion, and sometimes the complete extirpation, of many species. Historically, harvesting has already contributed to the demise of 55 terrestrial animal species (listed by the 2007 IUCN Red List of Threatened Species as Extinct or Extinct in the Wild), and 792 additional species are threatened with extinction from hunting (IUCN 2007). The Convention on International Trade in Endangered Species of Wild Fauna and Flora (CITES), an international agreement between governments that aims to ensure that international trade in specimens of wild animals and plants does not threaten their survival, has listed about 5,000 animals (including terrestrial, marine, and freshwater) for which trade regulations are in place (CITES 2008). Species extinctions preclude the provision of other present and future benefits. The large-bodied, slow-growing, species that are the most vulnerable to overharvesting (Bodmer et al. 1997) are often charismatic species that can attract tourists and funding, such as tigers, elephants, rhinos, and gorillas (Walpole & Leader-Williams 2002). Hunting also raises ethical and animal welfare issues (e.g. Gilbert 2000), and there is increasing resistance to the idea of killing animals for sport among urban residents in the industrialised world, as highlighted by the recent ban on fox hunting in the UK (Lindsey et al. 2006). Outbreaks in zoonotic diseases (ie shared between humans, domestic animals, and wildlife; e.g. SARS, West Nile virus), causing hundreds of billions of dollars of economic damage as well as mortality in humans, livestock and wildlife, have been attributed to the global wildlife trade (Karesh et al. 2005). 4.8.2 What are the overall trends in the provision of wild animal products? The Millennium Ecosystem Assessment (MEA 2005) considered that there is a medium to high certainty that wild food provisioning is declining, as natural habitats worldwide are under increasing pressure and wild animal populations are exploited at unsustainable levels (Wood et al. 2005). The most direct measure of trends in the provision of benefits from wild animal harvesting is derived from the flow of wildlife products, for example through statistics on consumption or trade. However, often these measures cannot be used to extrapolate future (even near- future) trends. Indeed, a steadily increasing trend in consumption or in market volumes may persist for some time despite depletion of the source populations. This may occur if there is increased harvesting effort (e.g. traps being replaced by guns), if harvesting extends to species not previously exploited, or if harvesting is expanding to new areas (Crookes et al. 2005). The higher the demand, the greater the economic incentive to source wildlife products further away, so a thriving market may persist for a substantial period. Ultimately, though, the widespread depletion of animal populations will inevitably result in declines in the supply of wildlife products. Overexploitation of wild animals is typically linked to human population growth, and consequently it is currently felt most in Asia, followed by Africa and by Latin America (Milner-Gulland et al. 2003). Depletion of animal populations is often associated with habitat loss, but may also occur in forests that otherwise appear to be intact (‘empty forest 107 syndrome’; Redford 1992). In Asia, a thriving trade of wild animal products for food, pets and traditional medicines results in widespread population declines and local extirpations (The World Bank 2005; Corlett 2007), with overexploitation being one of the most importants threat to some groups of species, such as amphibians (Stuart et al. 2004) and freshwater turtles and tortoises (van Dijk et al. 2000). In Vietnam, for example, 12 large vertebrates have been extirpated in the past 40 years largely because of hunting (Bennett & Rao 2002). In Africa, the scale of wild meat hunting has accelerated over the last few decades, reaching particularly worrying levels in Central and West Africa (e.g. Fa et al. 2006). For example, hunting has been identified as the main cause for the collapse in gorilla and common chimpanzee populations in western equatorial Africa (Walsh et al. 2003). In Latin America, where human population densities and resulting rates of exploitation are still considerably lower than in Africa (Fa et al. 2002), even subsistence hunting can substantially affect population densities and faunal community structure. In the Amazon, the aggregated biomass of the most sensitive species was reduced by half between non- hunted and light-to-moderately hunted sites, and reduced 11-fold between non-hunted and heavily hunted sites (Peres and Palacios 2007). Given that wild meat exploitation in tropical forests is often substantially higher than natural productivity (e.g., Fa et al. 2002) supplies in wild meat protein are predicted to decline (e.g. by 81% in the next 50 years, in the Congo Basin; Fa et al. 2003). On the other hand, game ranching (the maintenance of wild animals in defined areas delineated by fences) is expanding as a method of wild meat production, particularly in Sourthern Africa, although the main aim of these farms is to attract wildlife tourism or trophy hunting (Ntiamoa-Baidu 1997). As for trends in recreational hunting, participation in the US has declined by about 10% between 1996 and 2001 (USFWS 2007), while numbers of hunters in France have declined by about 3% per year (RSPB 2006). These declines may be explicable by increasing urbanisation (Heberlein et al. 2002), often associated with increasing animal welfare concerns. However, there is evidence that trophy hunting is expanding in Southern Africa and Tanzania, but stable or declining in Central and West Africa (Lindsey et al. 2007). 4.8.3 How is the provisioning of wild animal products affected by changes in wild nature? The supply of wild meat is largely determined by the ecosystem’s productivity of the targeted species. Often there is little consumer preference amongst species and they are mutually interchangeable as protein sources. Hunter effort varies spatially according to drivers such as market access and alternative livelihoods, and combines with overall animal abundance to determine supply. Hunting history affects current supply (Peres 1999). Large-bodied (average adults > 1 kg) primates, ungulates and rodents account for most of the wildlife biomass hunted by humans for food across the tropics, with ungulates typically making up by far most of the biomass. Robinson & Bennett (2004) reviewed data on the standing biomass of these taxonomic groups across a diversity of relatively undisturbed ecosystems, finding a peak for intermediate levels of rainfall. The increase in biomass with rainfall reflects a positive association with in primary productivity with rainfall, while the decline for highest rainfall levels may reflect a reduction in food availability, particularly for ungulates, with increasing forest density (as much of the plant biomass is in the form of inedible tree trunks and toxic leaves, with high proportion of the primary productivity high in the canopy; Robinson & Bennett [2004] and references therein). Accordingly, they found 108 that ungulates predominate in open habitats while primates are realtively more common in forested areas. Overall, their results indicate that mammalian standing biomass is low below 100 mm of annual rainfall, that grasslands with more than 500 mm can commonly support mammalian biomasses of between 15,000 and 20 000 kg/km?, and that the total mammalian biomass in tropical forests rarely exceeds 3,000 kg/km? (but see Fa & Peres 2001 for much higher values in Africa). Within a given habitat type, soil fertility can have a major effect on primary productivity and, therefore, on the standing biomass of mammal species; for example, Peres (2000) found that nutrient-rich floodplain forests consistently contained a greater game biomass than nutrient-poor unflooded forests. The sustainable production of wild meat is affected not only by the standing biomass but also by the species’ rate of natural increase. The percentage of standing biomass that can be harvested without resulting in population decline (the maximum sustainable offtake rate) is typically lower for long-lived species, estimated by Robinson & Bennett (2004) to exceed 50% for some rodent species, be around 20% for ungulate species, and usually under 5% for primates. Given that, as described above, ungulates and (to a lesser extent) rodents dominate in savannah habitats, while in forest primates play a more dominat role, the maximum sustainable offtake is likely to be higher in the former (maybe as much as 20% of the standing biomass) than in the latter (perhaps 10%; Robinson & Bennett 2004). A higher sustainable offtake combined with a larger standing biomass means that savannahs are typically more productive than closed forests, and therefore have a higher likelihood of being harvested sustainably (Robinson & Bennett 2004). While a wide diversity of species is usually harvested, the bulk of the wild meat biomass captured in any one location tends to be dominated by a few species (e.g. Fa et al. 2005, 2006). These, however, change in space and time. At least in South America, large species, such as tapirs, tend to be targeted by hunters first, with efforts turning to smaller species, such as squirrels or rats, as these disappear. This preference for larger-bodied species seems to be explained by a maximisation of return (kg of meat) per unit of hunting effort, rather than intrinsic cultural preferences for particular species (Jerozolimski & Peres 2003). Unfortunately, large species tend to have life-history traits (low reproductive rates, low population densities, long generation time, and long lifespans) that make them particularly vulnerable to overexploitation and less likely to recover once depleted (Bodmer et al. 1997; Jerozolimski & Peres 2003). As the most desirable species disappear, hunters’ selectivity declines and they tend to diversify their prey base (Jerozolimski & Peres 2003). Depletion of the large-bodied species may not result in a decline in wild meat supplies; theoretically, it may even result in an increase, as their reduced density allows faster-growing (more productive) species to dominate the ecosystem (density compensation; Peres & Dolman 2000; Jerozolimski & Peres 2003). Sustainable exploitation, in the sense of a future continuous supply of wild meat, may therefore become a more achievable aim in these conditions (Cowlishaw et al. 2005; but see Waite 2007), although this process cannot be considered sustainable in the ecological sense if it requires the extinction of a subset of species. Habitat loss inevitably results in the extirpation of habitat-dependent species and the wild food obtained from them. Nonetheless, a temporary increase in the supply of wild meat may be observed during the process of habitat loss, for example as new logging roads are opened, due to increased accessibility to wildlife. In the longer-term, though, supply is expected to decline as populations are depleted by the synergistic effects of hunting and habitat loss (Peres 2001). Habitat fragmentation reduces the habitat available to species 109 averse to the matrix and to edges. Peres (2001) calculated that the minimum fragment area needed to maintain a sustainable harvest of 46 hunted species in the Amazon (given levels of extraction documented to date) was as high as 2000 km?, with tapirs as the most spatially-demanding species. This is in sharp contrast with the observed size of fragments in the region of the deforestation frontier, with a modal area of less than | ha (Peres 2001). However, it is not clear that species diversity per se affects wild meat supply (see below). Fragmentation also isolates populations, thereby reducing or precluding colonisation of overharvested areas from nearby non-harvested or underharvested regions, disrupting the source-sink dynamics that help maintain the long-term sustainability of game harvest over large spatial scales (Novaro et al. 2000. 2005). This may explain why surprisingly few local extinctions have been reported in persistently overhunted (but finite) harvested areas within large, non-harvested, forest landscapes (Peres 2001; Ohl-Schacherer et al. 2007). Accordingly, the spatial configuration of harvested and non-harvested areas can affect the resilience of populations to harvest levels (McCullough 1996). Robinson & Bennett (2004) hypothesised that the conversion of primary forest to secondary habitats (such as forest—farm—fallow mosaics) may increase the supply of wild meat (see also Wilkie 2005). As discussed above, in dense forests much of the plant biomass is in the form of inedible tree trunks and toxic leaves, with high proportion of the primary productivity high in the canopy. Habitat disturbance opens the canopy, increasing the amount of browse and graze available for herbivorous animals such as ungulates, which can in principle increase the standing biomass of wild animals. However, disturbance is usually associated with increasing human presence (through opening of roads and logging), resulting in a decrease in the biomass of wildlife because of hunting and the introduction of competing livestock. On the other hand, hunting also shifts the composition of the faunal community, so that the biomass of large-bodied, slow reproducing species (including predators) declines, while the biomass of more adaptable, rapidly-reproducing, small- bodied species, might increase. Robinson & Bennett (2004) propose that the consequence of this shift is that even though overall biomass might decline, biomass production might, under certain circumstances, increase with the transition between primary forest and more disturbed habitats. This remains unresolved, with the studies reviewed by Robinson & Bennett (2004) providing some evidence in support of this currently open hypothesis. The effects may differ between continents depending on their respective faunal assemblages. In the Neotropics, most large vertebrates are primarily or partially frugivorous or granivorous (Robinson & Redford 1986b), and so a shift from fruit-based to browse-based plant food resources may not necessarily lead to an overall increase in the standing biomass and/or productivity of wild meat. Irrespective of whether secondary habitats are more or less productive than primary forest, they are certainly not empty of wildlife and they play an important role in wild meat supply in many areas (Daily et al. 2003; Naughton-Treves et al. 2003). In theory, a wider diversity of targeted species could directly improve the supply of wild meat in two possible ways: - By increasing the secondary productivity of the ecosystem, if different species explore ecological resources in a complementary way. This effect has been predicted theoretically and demonstrated in experiments with simplified plant systems (Hooper et al. 2005) but we found no evidence that it may happen for the provision of wild meat. 110 - Through a portfolio effect, by providing a diversity of options to hunters. Variety does not seem to be a major factor from the consumer’s perspective, with the most abundant species often being the most preferred (but see East et al. 2005). A diversity of species could also allow hunters to switch to other species as some become too scarce. Indeed this is the case, but species disappearance tends to follow a nested pattern, with the most sensitive only being present in species-rich systems and the most tolerant species dominating species-poor systems. As discussed below, we found no evidence that the latter are less productive than the former. Species diversity may nonetheless affect the provision of wild meat indirectly, by influencing ecosystem composition, spatial structure, and ultimately resilience. The species that tend to be the most affected by hunting often play key roles in maintaining tropical forests, performing ecological functions (grazing, browsing, trampling, seed dispersal and excavation) that are disproportionately large relative to their total numbers (Robinson et al. 1999; Wilkie & Lee 2004). In particular, depletion and extinction of frugivorous species seems to have a significant effect on seed dispersal and consequently on the recruitment of particular plant species (Roldan & Simonetti 2001). Large-bodied primates (woolly monkeys and spider monkeys), in the Amazon are key to the recruitment of many large- seeded tree and liana species (Peres & Palacio 2007), while tapirs are key to the recruitment of a large-seeded palm in the Amazon (Fragoso 1997). In the Congo Basin, forest elephants are keystone species, ecosystem engineers, and the bulk of mammalian biomass where they occur. The decline of these species (all of them favoured by hunters) may therefore result in long-term changes in the forest communities, both in terms of floristic compositional and of spatial structure (Stoner et al 2007). The abundance and diversity of other (non-targeted) species can theoretically affect wild meat production through either bottom-up or top-down processes (Sinclair & Krebs 2002). Bottom-up effects are more straightforward in that a higher abundance of food species can improve animal productivity. Diversity is also likely to have a positive effect, increasing the stability of food provision across the year. Top-down effects are more contentious. The effect of a higher number of predators is the decline in prey abundance, including many species (such as ungulates, rodents, primates, and squirrels) that are the basis of wild meat supply. And indeed Michalski & Peres (2007) found that a decline in predators leads to an increase in wild meat supply through predator release. However, a diversity of predator species may have a positive effect, if top predators prey on other predator species (mesopredators). Theoretical studies (Palomares et al. 1995) and some empirical results (Crooks & Soulé 1999; Terborgh et al. 2001) provide some evidence for this effect. Recreational hunting (and trophy hunting in particular) is more selective in terms of the species targeted, and so it is likely to be more affected than harvesting of wild meat by the loss of particular species. Relationship between habitat area and wild meat provision The diversity of harvested species is predicted to increase with area (MacArthur & Wilson 1967), and indeed Michalski & Peres (2007) found a classic species-area relationship (linear with the logarithm of area) for mammals species in a fragmented landscape in the southern Amazon (Figure 15a). Peres (2001) predicted a non-linear relationship between harvested vertebrates and logarithm of forest fragment area: small fragments (<20 ha) retain practically no species (with or without hunting pressure); larger areas are required to retain the same number of species when hunting pressure is higher (e.g. 90% of the species 111 retained in fragments of 3000 ha in the absence of hunting, but required about 11,000 ha if exposed to moderate to heavy hunting); and areas as high as 2000 km” required to maintain the most spatially-demanding species (tapirs) (Figure 15a). These studies also found that as area declines, a non-random, nested simplification of fauna takes place, with large slow- growing species, or those that live in large wide-ranging groups that occur at low density, disappearing first. The relationship between area and species richness is not particularly informative, though, as we found no evidence that species richness of harvested species drives wild meat provision. The relevant relationship is one between area and wild meat productivity (biomass that can be harvested, a product of standing biomass and the maximum sustainable offtake rate). We did not find any study which explicitly investigated this relationship. In principle, the relationship between wild meat productivity and area can be: a) Neutral: productivity per unit of area is the same irrespective of whether that unit of area belongs to a small or a large habitat fragment. This is plausible if we assume that animal productivity reflects primary productivity, and that this remains the same irrespective of fragment size. In this case, wild meat production increases linearly with area (Figure 15b, blue line). b) Positive: larger fragments have larger productivities per unit of area. This can be the case if the reduction in species richness for declining fragment area translates into a decline in overall wild meat productivity (for example, if it means that some ecological niches become empty, ‘wasting’ some of the primary productivity). In this case, productivity increases faster than area, at least until the point where area becomes large enough to maintain the complete faunal assemblage (Peres 2001 found this to be about 200,000 ha in Amazonian forests) (Figure 15b, purple line). c) Negative: smaller fragments have larger productivities per unit of area. This is plausible if as faunal assemblages become more simplified for smaller areas, they become dominated by species that are on average more productive (e.g., if slow- reproducing species are replaced with fast-growing species). Predator release (Terborgh et al. 2001) may also result in higher productivity in smaller habitat fragments. In this case, productivity increases slower than area, again at least until the point where area becomes large enough to maintain the complete faunal assemblage (Figure 15b, green line). Edge effects may also affect the shape of the relationship between area and the productivity of wild meat: d) If the matrix habitat is unfavourable (e.g. a population sink due to increased mortality) it is expected that smaller fragments (with more pronounced edge effects) will have lower productivities per unit of area, and productivity will increase faster than area while the fragment is small enough to have all of its area affected by edge effects (this will vary with the distance at which edge effect are felt; for 1 km deep edge effect in a circular fragment, the area would be 314 ha) (this produces the same shape as for Figure 15b, purple line, but the threshold would be a different area). e) Ifthe matrix habitat is favourable (e.g. by providing additional sources of nutrition) it is expected that smaller fragments (with more pronounced edge effects) will have higher productivities per unit of area, and again productivity will increases slower 112 than area while the fragment is small enough to have all of its area affected by edge effects (Figure 15b, green line). b a) 4 ) 40 4 SS = o — ee OE 2 2 = s g 1S) ne} i coe 9 ® fom oO ® o ‘ Qa @O ‘ Oo £ : U : 2 1 10 100 1000 10000 100000 forest fragment size (ha) forest fragment size (ha) Figure 15. a) Relationship between forest fragment size (note the logarithmic scale) and the diversity of mammal species (from Michalski and Peres 2007). b) Possible shape of the relationship between wild meat productivity (in biomass per year) and forest fragment size (see text for details). Michalski & Peres (2007) found a negative relationship between forest-patch area and two measures of the aggregate abundance of mammal species, and no variation for a third measure. However, overall abundance (number of individuals per unit area) is not an ideal measure of wild meat supply because it does not account for variation in body size across species. Indeed, if larger species are dominant in larger habitat patches, then a negative relationship is expected between area and abundance, even if standing biomass remains constant. Michalski & Peres (2007) found that larger species (e.g., tapir, giant armadillo) only occurred in the largest fragments. On the other hand, larger-bodied species tend to reproduce slower, and so for the same standing biomass productivity may be higher for smaller fragments if they are dominated by small, fast-reproducing species (supporting hypothesis c, above). Michalski & Peres (2007) also found that for several species, there was a hyperabundance in small patches. They explained this as possibly resulting from: subsidies by matrix habitats (e.g. presence of exotic fruit trees; supporting hypothesis e); the outcome of density compensation (increase in the abundance of one species as competition with other species declines; supporting hypothesis a); or the outcome of release from top predators (increase in abundance due to predator removal; supporting hypothesis e). In summary, wild meat production likely increases linearly with area, but a threshold habitat size may exist below which the relationship is non-linear. 4.8.4 What are the main threats to the provisioning of wild animal products? Overexploitation is the main threat to the provision of benefits from wild animal harvesting, frequently resulting in the exhaustion of the source populations. Hunting pressure in 113 tropical forests has risen dramatically in recent decades, concomitantly with forest loss, increasing human population, greater access for hunters and traders to remaining forests as a result of road building and forest fragmentation, the use of efficient modern hunting technologies (especially firearms and wire snares), loss of traditional hunting controls, and greatly increased commercialization of hunting (Milner-Gulland et al. 2003). Hunter access to inaccessible forests is exacerbated by commercial logging operations: these create an extensive network of roads linked to the national road system, and the trucks that travel them become conduits for a vast commercial trade in wild meat, transporting it from remote areas to towns for sale (Robinson et al 1999). Overexploitation can lead to the local or global global extinction of particular species, even if these have become so scarce that it would be uneconomic to pursue them individually. This is because hunters pursuing other more common species will still kill scarce species encountered (e.g. Clayton et al. 1997). Even the collection of other non-timber forest products (e.g. Brazil nuts and forest vines) may exacerbate the possibility of this “piggyback extinction’ to take place, by lowering the opportunity costs of hunting (L. Parry, unpublished data). Increased demand for wild meat may also result from a lack of availability of alternative protein sources (Rowcliffe et al. 2005). Brashares et al. (2004) found that a declines in fish supply in West Africa were associated with increased hunting in nature reserves and ultimately the sharp declines in the biomass of 41 wildlife species. The introduction of livestock may reduce wild meat supplies in savannah systems, if the domestic species compete with the wild ones (Robinson & Bennett 2004), but on the other hand it may reduce demand for wild meat (and therefore prevent overharvesting) if it provides a more readily accessible source of protein (Bennett 2002). Outbreaks of human commensal species (for example baboons) may cause the decline of the harvested species and prevent them from recovering even if the hunting activities are subsequently reduced or ceased (J. Brashares, personal observation). Increased contact between people, livestock and wildlife increase the likelihood of disease outbreaks in any of these groups (Karesh et al. 2005; Rizkalla 2007), therefore posing risks not only to wild meat provision but also to public health and to economic activities. Social and economic disruption can affect the provision of wild meat by creating the conditions for overharvesting, for example by displacing people and promoting illegal trade. They can also affect recreational hunting, particularly international, by reducing expenditure and willingness to travel. 4.8.5 Are abrupt changes likely in the provision of wild animal products? At the single-species level, overexploitation can lead to sudden transitions from large harvests to population collapses (Barnes 2002). We predict that benefits derived from a single or just a few species (e.g. a particular type of medicine or fur) will undergo sudden collapses as well. However, benefits from wild meat production, often based on harvesting a broad set of species, boom-and-bust trajectories of individual species are likely to be obscured as hunting effort extends to other species. Therefore, changes in overall wild meat supply are likely to be gradual, except when faunal communities become very simplified and hence a sudden collapse may be expected if the last few species all decline abruptly. 114 As described above, large-bodied species that are most sensitive to overharvesting play key roles in maintaining the habitats themselves (e.g. through seed dispersal), and so the disappearance of these species will result in a permanent change in ecosystem structure and composition (Fragoso 1997; Robinson 1999; Fragoso et al. 2003; Peres & Palacio 2007), which in turn may influence habitat quality for other species, thereby affecting the provision of wild meat in the long-term. This process, which is irreversible once species are permanently lost, may take decades or even centuries for its effects to become obvious. Overall, the key factors that are more likely to either result in or prevent a very quick depletion of wild meat supply are socioeconomic in nature, rather than biological. These include rights and access to the resource, management of the resource, availability of alternative protein sources, markets, economics of harvest, and incentives to use species sustainably. Recreational hunting, and in particular international trophy hunting, is also particularly sensitive to socioeconomic changes, likely to collapse as a result of political instability in a region (e.g. war, civil disruption). While its reliance on a smaller set of species could in principle make this activity more prone to disintegrate, in practice it is unlikely in cases where the exploited species and populations are actively managed. Overall, we predict that there is a very high probability that individual populations of harvested species (particularly those harvested for subsistence or commercially) will collapse in the near future. However, there is a low probability that this will translate into a sudden collapse in the provision of wild meat. Socioeconomic changes are the most likely factors that could result in a sudden change in the supply of wild meat or in recreational hunting. 4.8.6 Can we quantify and map the global provision of this benefit/process and how it might change? We aim to obtain a method or model for generating, from a given map of land cover, a surface of the estimated maximum sustainable production of wild meat. Given two maps (two ‘states of the world’) we would then be able to estimate the difference in the provision of wild meat between one state and the other. The economic value of this difference would then be calculated using information on demand, having into account that value (prices) are a function of both supply and demand. A first cut model would calculate local production values (kg/km*/y), derived from empirical data for particular sites, and then generalised from local production estimates to larger-scale patterns, in order to create a map of potential wild meat production for the regions analysed, under current circumstances. This map would be heavily biased by the fact that production is dynamic, with different population sizes producing different levels of offtake at different hunting pressures; in order to generalise robustly about maximum sustainable yield values, a time series of population size and offtake estimates is needed from which to calculate sustainable offtake rates - of which there are very few in tropical forest areas. 115 We suggest following Robinson & Bennett (2004) in focusing on the extraction of large (> lkg) mammals, as these dominate supply of wild meat as well as available data on species densities. We also suggest restricting the analysis to Central and South America, sub- Saharan Africa, and tropical Asia and Australasia, as these are the main regions where wild food is particularly important for human wellbeing. A first approach for calculating local production rates is the widely used Robinson and Redford (1991) model (e.g. Peres 2001; Fa et al. 2002; Robinson & Bennett 2004). This predicts the maximum sustainable wild meat offtake for each species in a given habitat, that is, the percentage of standing biomass that can be harvested without resulting in population decline. Calculating this offtake requires information on: - The carrying capacity of the habitat for each species, typically measured as the density (number of adult individuals per unit area) of the species in unexploited areas. This can also be estimated from empirical relationships between density, diet and body size. - The maximum intrinsic rate of natural increase (7max) for each species, the maximum rate of increase that a population can achieve under natural conditions without significant intraspecific competition (Robinson and Redford 1986). This is very difficult to measure directly (Milner-Gulland & Akgakaya 2001). Fast-reproducing species such as ungulates and rodents have higher intrinsic rates of natural increase than primates. - A correction factor to account for natural rates of mortality (generally varying between 0.2 and 0.6), that is lower for longer-lived species. Both rmax and carrying capacity can be roughly estimated from their allometric relationships with body size, corrected by taxonomic group and geographic location (Rowcliffe et al. 2003). Using Robinson and Redford’s (1991) model, the maximum sustainable offtake rate results in percentage values that are typically lower for long-lived species: annual offtake rates for some rodent species exceed 50% of the standing biomass; are generally lower, but frequently above exceed 20%, for ungulates; and are very low, usually under 5% of standing biomass, for primates (Robinson & Bennett 2004). The overall production of a given area (wild meat biomass that can be extracted per year) therefore depends on the particular set of species in the community, as their life history traits, body size and relative abundance affect the final result. For example, using this method Fa et al. (2002) calculated a production of 1,111 kg/km*/year for the Congo Basin and of 488 kg/km’/year for the Amazon. There are, however, important and fundamental limitations with this approach and we strongly recommend that a more suitable model (or ideally, a battery of different models) is used to calculate sustainable production rates (Milner-Gulland & Akgakaya 2001). The model by Robinson & Redford (2001) assumes that the aim is to maintain populations at their original size, yet harvest may bring populations to any lower population size and still be sustainable if the population remains stable. These lower population sizes may be more desirable to improve long-term profitability or to maintain pest populations at low 116 densities. Deciding at which population size the “maximum sustainable harvest” can be taken is therefore a priority. One option is to assume that all target populations should remain above some percentage of their carrying capacity. The tradeoff is then between maximising production (at a population size of 50% of carrying capacity, under the simplest assumptions) and reducing the risk of population extirpation in the face of environmental stochasticity (at a population size greater than 50% of carrying capacity). A number of authors have suggested 75% of K as a suitable precautionary reference point (Roughgarden & Fraser 1996). Having decided the target population size, one needs to estimate how much wild meat could be produced sustainably. This requires understanding the rate at which populations can grow. Robinson & Redford (2001) use the maximum intrinsic rate of natural increase, but this overestimates the rate at which populations can grow unless they are at a very low levels. For higher population sizes, the growth rate tends to zero as carrying capacity is approached, due to intra-specific competition (density dependence). Population growth rates should therefore be scaled according to population size (i.e. how close it is from maximum carrying capacity). A first approach for a growth model incorporating density- dependence is the logistic function (Milner-Gulland & Akcakaya 2001). Assuming logistic growth, the maximum sustainable production at 50% of carrying capacity is “4 of the intrinsic rate of increase multiplied by the carrying capacity. However this is a single- species model which would be unsuitable for complex multi-species systems such as bushmeat harvesting. There has not yet been any work done to calculate a maximum sustainable production rate for such systems. A substantial difficulty with the application of any of these models is for areas that are already being harvested, which is the case for the vast majority of tropical forest areas. Here, the current population densities cannot be considered to be the population size at carrying capacity. Measures of offtake need to include not only reported captures but also collateral mortality — instances in which animals are fatally wounded by hunters yet not recovered. In one game harvest study in the Amazon, modern colonists hunting with shotguns caused an estimated 9% collateral mortality (L. Parry, unpublished data), while wastage rates in snares in Africa have been estimated at up to 28% (Muchaal & Ngandjui 1999). Finally, estimates of productivity should incorporate uncertainty in parameter estimates, and appropriate methods are becoming available for doing so (Milner-Gulland & Akcakaya 2001). Moving from local estimates of population productivity for individual species to large-scale surfaces of overall wild meat productivity could take two possible routes. One could generalise first on a species-by-species basis, using information on habitat suitability to infer changes in species density, to produce a map of potential productivity for each species. These could then be added to produce overall productivity maps. A second route is to aggregate the productivity of all species at a given site to obtain an overall estimate of wild meat productivity at current population sizes and harvest rates, and then generalise those across space based on bioclimatic and land cover data. We recommend the second route, as the former would require much finer, species-by-species, data that are unlikely to be available, as well as making untested assumptions about the interspecific interactions between species. 117 High quality empirical data would inform better models, as more explanatory variables can be integrated to explain local production. For example, the incorporation of adequate soil fertility data will improve production models since nutrient-rich environments have a greater game biomass than nutrient-poor ones; as shown by Peres (2000) for floodplain and unflooded forests in the Amazon. Additionally, livestock densities will affect the density of wildlife through competition, thus livestock distribution maps can be incorporated, as well as information derived from the Human Footprint Map on the degree of human disturbance, fragmentation and alteration (Sanderson et al. 2002). Data limitations mean that only a very crude model can be generated. Existing data on local species composition and densities of individual species (e.g. Fa & Purvis 1997; Peres 2000; Robinson & Bennett 2004) are still very patchy, although improving. Studies also integrating accurate measures of offtake are even scarcer, as are studies for secondary habitats in general (including secondary forest and agricultural land). We suspect there is a vast wealth of data on the productivity of different habitats (e.g., forests, wetlands, mountain areas) for species of birds and mammals harvested through recreational hunting in Europe and North America. This is expected given the degree to which hunting is generally managed in these regions. However, we do not know if such data are concentrated or dispersed. If the former, models of productivity for recreational hunting would not only be possible but also substantially better than the models for wild meat harvesting in tropical regions, as they would rely on better data and apply to much less complex communities (with fewer species). Modelling of recreational hunting of migratory species (such as ducks) would however pose specific challenges. As for international trophy hunting, we are under the impression it mostly takes place in private land or concessions that are actively managed for the target species. Sustainable harvests are likely to be ensured by managers, who not only control harvest directly but also enhance the habitat to increase species’ productivities. The economic value that can be gained from this activity is therefore determined much more by the capacity of operators to attract tourists than by biological constraints. In this sense it is much more like international tourism (section Error! Reference source not found.), and should be modelled and valued accordingly. Adequacy of scenarios The key information that the scenarios need to produce is maps of landcover. Other layers likely to be important include: human density, transport routes, livestock density, and the bioclimatic variables used to model variation in productivity across space. 118 What can be done in Phase 2? At what cost? By whom*? A first-cut, crude, pan-tropical, model is possible given the available analytical tools, and could be developed within one year given 24 researcher-months. Several research groups would be well positioned to do this work (see key experts below). The quality of the model would rely crucially on the underlying data. Maximising the quality and quantity of the data that could be made available would be best done by establishing collaborations across institutions working in different areas. Our recommendation is therefore that this work builds from a consortium, brought together through two NCEAS-style workshops. We suspect that a model of recreational (domestic) hunting is possible for Europe and North-America, if data on game productivity are concentrated in particular institutions (perhaps CIC, see key resources). We recommend investing some time in pursuing this possibility as if such data repositories exist, creating a model would be both feasible and economically relevant. 4.8.7 Insights for economic valuation Published economic valuations of wild meat harvesting are typically based on the current rates of extraction (e.g., Gram 2001; de Merode et al. 2004), yet these are known to be frequently unsustainable (e.g., Fa et al. 2002). Spatially-explicit models of the potential value of habitats for bushmeat production cannot therefore be built from studies on current values of wild meat extraction. Instead, biophysical models of sustainable production could be generated first, based on very crude models built using allometry to estimate rmax and carrying capacity, and the few studies that exist on population sizes in unhunted and lightly hunted areas. An estimate of the associated uncertainty could also be included. These predictions could be compared with spatially explicit maps of current bushmeat production, drawn from broad-scale surveys (such as are being developed by the Bushmeat Crisis Task Force, www.bushmeat.org). Crude extrapolation from existing data to unsurveyed areas on the basis of human population density, vegetation type and transport routes would be possible, and these could highlight areas in which hunting is most likely to be highly unsustainable. Current economic values could be obtained by overlaying data on current prices and consumption. However because supply and demand are interlinked, and dependent on availability of both bushmeat and alternatives (both for consumers and producers), as well as income and tastes, use of these same data to estimate the potential economic value of sustainable bushmeat harvest would involve heroic assumptions. * This is our recommendation for Phase II, based on the results of this review. It does not commit the leaders of Phase II to follow it, and it does not commit the recommended research group to actually do such work, 119 Further extrapolation to alternative scenarios might be possible, but only if an understanding of the underlying drivers of changes in bushmeat hunting pressure and value were obtained. This has not yet been done even at the regional or local level, although there are datasets available for which this could be done. 4.8.8 Some key resources Wild food: e WCS Hunting and Wildlife Trade Program, led by Liz Bennett. Also at WCS are John Robinson and Kent Redford. e Imperial College Conservation Science group, led by E.J. Milner-Gulland. e Durrell Wildlife Conservation Trust, with John Fa as director of Conservation. e Bushmeat Research Programme at the Zoological Society of London, led by Marcus Rowcliffe and Guy Cowlishaw. e Justin Brashares group at University of California Berkeley, including Karen Weinbaum. e Carlos Peres group at University of East Anglia, including Luke Parry. e Washington-based Bushmeat Crisis Task Force, led by Heather Eaves. e Robin Naidoo, Taylor Ricketts and Neil Burgess at WWF-US. Recreational hunting: e FACE-Europe, a European federation of national hunters’ associations of the Member States of the European Union and other Council of Europe countries (representing the interests of some 7 million European hunters). e Ducks Unlimited, a North-American NGO focused on the conservation of waterfowl and wetlands for hunting purposes. e CIC — International Coucil for Game and Wildlife Conservation a Hungary-based non-governemental organisation focused on the management of wild-living resources. 4.8.9 Participants Authors James J.J. Waters, Ana Rodrigues, Andrew Balmford (University of Cambridge, UK) E.J. Milner-Gulland (Imperial College London, UK) John Fa (Durrell Wildlife Conservation Trust, UK) 120 o AT ty [omni che rea + 5 hi im ’ hich i hide am, aT A pes Oh yar ele i, sh HEMI . eit sien era : o Oe 3 - ws ea ; ® foe page? ORD ah i nite by Py a re va ; i iy hy i i ui di inate) as oat one % art oO RL Le Com A Muy eye ft any hod was tenis wp «=. © «@ “oe , ih, Cm te ee | a a oO wyotin ee ews wel tel eae te ] a Reta ‘Dt cag La ie aed, Pin. . i . Ai ee et atin The aon uti met i ; ! Pvt a Baie. a bag nerve la ay ae i oT inn . ech 7} (lynne it & Sur op NY a il FISTS Die, WOT sO ca dor aie hie » canh. ay th Rema Wil scinesitsy icy ee “the yew. teal U ae fle: rage wheal cu, Brea irny on et TWyeh), — ond ar oT ; ie Pavntey Te reese Fin) svi ye dg a ‘tall pemanyeean on =a) wot waper A a han A comida eal rilew Toe Peaamen Spa. rs ; PRE ere Hemline: teaduater th! hailey’ bor penvietong ipa i ae Tl a, ‘Miele Tiere Seedy aus ee) le Wehbe npr VAT oad oe mete tush “ihe: Carve suthek thin ope ow Mey «elle td Be aiid BranFearigs (peng Tt, OUD iI iy; mane iam A mare Z| Mit Gober wn hi otha np ! vey = days ue tp dd act quate ia Lig fe vi Boch ang Ollie? 4 eas PROVO ES watt sak, pihiat Wrgl | Sy tis Ay fecsyh (ua. a! Wal a pe Ane bawe prvi a bl ‘@ ie u IS i f j waier eit aapria i prio hone & +a rah Oy MeziGic ene withrdiweale TH ' ee pulieti uit Calvi rh ; sabes Le rane ofem " aioe tie grates a 4 beri ive iehqud Cato etion Pine Reivers We Det 0 eAligs rw oe he ci enon pie Tue . tha: Karen Weinbaum, Justin Brashares (University of California Berkeley, USA). Luke Parry (University of East Anglia, UK) Marcus Rowcliffe (Institute of Zoology, London, UK) Contributors John Robinson Acknowledgements Guy Cowlishaw and Elizabeth Bennett provided references and contacts. 4.9 Fresh water provision, regulation, and purification This section is a developed overview based on a literature review and expert consultation. It greatly benefited from expert contributions (see below) but it has not been reviewed by experts in this field. 4.9.1 Why is this process important for human wellbeing? Fresh water is fundamental to the survival of every single human, for direct consumption and to maintain an adequate food supply (Pimentel et al. 2004). Although water covers 71% of the planet surface, globally freshwater is a scarce resource: only 2.5 % of water is fresh, and nearly three quarters of that is frozen, and most of the remainder is present as soil moisture or lies deep in the ground. The principal sources of fresh water that are available to society reside in lakes, rivers, wetlands, and shallow groundwater aquifers—all of which make up but a tiny fraction (tenths of 1%) of all water on Earth. This amount is regularly renewed by rainfall and snowfall and is therefore available on a sustainable basis (V6résmarty et al. 2005; Arthurton et al. 2007). Furthermore, water is unevenly distributed across the planet’s surface, with abundant supplies in regions such as the wet tropics and absolute water scarcity across the desert belts and in the rain shadow of mountains (V6résmarty et al. 2005). Fresh water is essential first and foremost for drinking. Each person needs only 20 to 50 liters of water free of harmful contaminants each day for drinking and personal hygiene to survive, yet there remain substantial challenges to providing this basic service to large segments of the human population. Half of the urban population in Africa, Asia, and Latin America and the Caribbean suffers from one or more diseases associated with inadequate water and sanitation (V6rdsmarty et al. 2005). Globally, more than 1.1 billion people lack adequate access to clean water and 2.6 billion people — half of the developing world’s population — lack access to adequate sanitation. Every year some 1.8 million children die as a result of diarrhoea and other diseases caused by unclean water and poor sanitation. The world’s governments set as a global target under the Millennium Development Goals to halve by 2015 the proportion of world population without sustainable access to safe drinking water and basic sanitation (Goal 7, target 10) (UNDP 2006). Household water requirements represent a very small fraction of water use. Water use is dominated by agricultural withdrawals (70% of all use and 85% of consumptive use), including livestock production, followed by industrial and only then domestic applications. For example, approximately 1000 litres of water are required to produce | kilogram of cereal grain, and 43,000 litres to produce | kilogram of beef (Pimentel et al. 2004). Fresh water is also fundamental for inland fisheries, including aquaculture (section 4.7), for recreation (section Error! Reference source not found.), for the production of hydroelectric energy and for transport (waterways). Indirectly, freshwater water flows also contribute to marine fisheries production (section 4.6) (Table 3). Over the last few centuries, global water use increased roughly exponentially, and substantially quicker than population growth, as a result of both population growth and economic development, and in particular the expansion of irrigated agriculture (V6résmarty et al. 2005). Global water withdrawals now total ~3900 km’ yr, or ~10% of the total global renewable resource, and the consumptive use of water (not returned to the watershed) is estimated to be ~1800 to 2300 km* yr! (Foley et al. 2005). Water use efficiency is now improving, with per capita use rates dropping as of around 1980 from about 700 to 600 cubic meters per year (V6résmarty et al. 2005). The aggregate global withdrawal continues to increase, though, with water withdrawals predicted to increase, by 2050, by 50% in developing countries, and 18% in developed countries (UNESCO-W WAP 2006). While demand is increasing, freshwater resources are under threat by unsustainable water extraction and pollution. Particularly serious is the uncontrolled rate of water withdrawal from aquifers, which in many regions (especially arid ones) is significantly faster than the natural rate of recharge. This mining of groundwater reserves poses a serious threat to water supplies in world agricultural regions, especially for irrigation (Postel 1999). Increases in pollution of surface and groundwater resources not only pose a threat to public and environmental health but also contribute to the high costs of water treatment, thus further limiting the availability of water for use (Postel 1999). Ecosystems play important roles in the hydrological cycle, contributing to water provision (quantity, defined as total water yield), regulation (timing, the seasonal distribution of flows) and purification (quality, including biological purity as well as sediment load) (Dudley & Stolton 2003; Bruijnzeel 2004; Brauman et al. 2007). Different benefits are affected differently by each of these aspects of water regulation (Table 3). The production of these benefits may be synergistic or competitive, with water management for one benefit possibly affecting others (e.g., improvement of water quality for drinking purposes through reforestation resulting in a decline in water quantity for hydroelectric production). And use may be extractive (water is removed from the system), in-situ (when the benefits are enjoyed without requiring water to be extracted), and indirect (when water contributes to the benefit provision only indirectly). Table 3. Fresh water provision and regulation contributes to the production of a diversity of benefits. Water provision (quality), regulation (timing) and purification (quality) are not equally important for each of these benefits (estimates of relative importance: + somewhat important, ++ important, +++ very important). Type ofuse _—_ Benefits Water Water Water quantity _ timing quality Extractive Domestic water consumption + lata ctataaha use MONO PADD BAU OM orwell te en att la a aliivestocksproduction _ 07 ee a eh ee Industrial use ite sti + Ine Seeulase | tulancittshenes,(ineludin aquaculture) Sit Sec ta abiydnocleeniernerey production ee Gass a ae Transport: waterways +++ +++ + Recreation (sport, tourism) ++ ++ +4++ Indirect use Marine fisheries ++ 4+ aaere 4.9.2 What are the overall trends in fresh water provision, regulation, and purification? The Millennium Ecosystem Assessment (MEA 2005) considered that there is a medium to high certainty that freshwater provisioning, regulation and purification has been degraded in the recent past, caused by unsustainable use for drinking, industry and irrigation, and that there has been a decline in water purification services. Dams are increasing our capacity to use hydroelectric energy and water regulation services have changed in a variable way, depending on ecosystem change and location. 4.9.3 How is the provision, regulation, and purification of fresh water affected by changes in wild nature? Water provision (quantity) In reviewing what is known about the impact of biodiversity on water quantity, we rely extensively on the exhaustive review Bruijnzeel (2004). Here we summarise key points about the influence of forest cover on regional climate (rainfall) and total water yield; in later sections we précis what is known about its effects on the timing of water yields, and on erosion and sedimentation. The quantity of water delivered from a watershed is conventionally measured only as surface water output and reported as mean annual watershed yield. Ecosystems, however, affect the available quantity of both surface and groundwater (Brauman et al. 2007). Vegetation, and forests in particular, significantly affect the quantity of water circulating in a watershed, although not necessarily in the way that is popularly perceived. Dealing first with precipitation, while it is commonly assumed that forests generate rainfall, this is not straightforward to test, particularly in tropical regions, because rainfall can be highly variable in space and time (Bruijnzeel 2004). The rationale for this effect is that the higher evapotranspiration and greater aerodynamic roughness of forests (compared to pasture and agricultural crops) leads to increased atmospheric humidity and moisture convergence, and thus to higher probabilities of cloud formation and rainfall generation. Early demonstrations of this effect have however been questioned, with the suggestion that observed higher rainfall in forested areas was an artifact of orographic effects (remaining forests tend to be in uplands where clouds are forming because of atmospheric cooling of rising air) or of differences in exposure of measurement instruments to wind and rain. Bruijnzeel’s (2004) review concluded that there is increasing evidence that large-scale land use conversion affect cloud formation and rainfall patterns, although the effects may be small and they depend on the particular land use conversion. For example, simulation predicts large-scale Amazonian forest conversion to pastureland would result in a mere 7% reduction in annual rainfall; conversion between primary forest and secondary forest or even some crops (e.g. tea plantations) is expected to have little effect on levels of evapotranspiration; and forest conversion to rice paddies is expected to increase cloud formation and rainfall (Bruijnzeel 2004 and references therein). It is well-established that, in coastal or montane regions subject to wind-driven fog or clouds (cloud forests), the presence of tall vegetation can significantly increase the amount of water reaching the forest floor as canopy drip (Bruijnzeel 2004). It has been also been suggested that deforestation of upwind lowland forests can reduce cloud cover on adjacent upland cloud forests, affecting moisture interception by the canopy (Lawton et al. 2001). In contrast with popular belief, it is well-established that the dominant impact of vegetation, and of forests in particular, is one of reducing the surface water output (which ultimately forms rivers), thereby reducing watersheds yield (Dudley & Stolton 2003; Bruijnzeel 2004; Brauman et al. 2007). This effect is due to evapotranspiration, whereby plants draw water from the soil into the atmosphere as part of photosynthesis. Trees generally use more water than other plants, as their roots allow them to absorb water from deeper in the soil. Also, being aerodynamically rougher than short, smooth vegetation, they increase the efficiency of gas exchanges with the atmosphere (Brauman et al. 2007). Under mature tropical rain forest, typically 80-95% of incident rainfall infiltrates into the soil, of which ca. 1000 mm per year is transpired again by the trees when soil moisture is not limiting, whereas the remainder is used to sustain streamflow (Bruijnzeel 2004). Bruijnzeel (2004) reviewed the results of controlled experiments comparing water yield before and after deforestation, finding that in all cases the removal of more than 33% of forest cover resulted in significant increases in overall annual streamflow during the first 3 years, with changes in water yield roughly proportional to the fraction of biomass removed. As forest regenerates, or is replaced by forest plantations, water yields tend to decline again to pre-clearing levels, a process which happens faster in tropical regions (where vegetation grows more quickly) than in temperate areas. In contrast, conversion to cropland or pastureland may result in permanent increases in annual water yields. This reflects not only the diminished capacity of short vegetation to intercept and evaporate rainfall (given its lower aerodynamic roughness and generally smaller leaf area) but also its reduced capacity to extract water from deeper soil layers during periods of drought (because of more limited rooting depth) (Bruijnzeel 2004 and references therein). While there are no reported declines in annual streamflow totals following lowland tropical forest removal, from the opposite appears to hold for the conversion of montane cloud forests to cropland or pastureland. This is because evapotranspiration in generally low in cloud forests, while cloud forest-clearing typically results a reduction in flows generated by cloud and fog interception (Bruijnzeel 2004; Postel & Thompson 2005). Water requirements are different for different plant species, so species composition of vegetation may affect water yields. Young and invasive plants generally have disproportionately large impacts on water quantity because vigorously growing vegetation tends to use more water than mature vegetation. In arid areas, adaptations such as dry- season senescence of native vegetation may limit its water use, whereas an introduced species that lacked these traits would consume water over longer periods during the year (Brauman et al. 2007 and references therein). In South Africa, for example, the spread of non-native eucalyptus, pine, black wattle, and other thirsty trees into the native fynbos (shrubland) watersheds of the Western Cape is having negative consequences for water supplies, as their evapotranspiration requirements greatly exceeds that of the low-stature and drought-tolerant fynbos vegetation (van Wilgen et al. 1996). 125 Lakes and freshwater wetlands store about 50 times as much water as rivers (Vérésmarty et al. 2005) and, if well-managed, provide a stable water supply for domestic, agricultural, and industrial use (Finlayson et al. 2005). Some wetlands provide an important role in aquifer recharge, therefore contributing to groundwater supply. However, some wetlands exist because they overlie impermeable soils or rocks and there is, therefore, little or no interaction with groundwater (Fenlayson et al. 2005). Wetlands evaporate more water than other land types (including grassland, forests, and arable land) and therefore reduce annual water yields (Bullock & Acreman 2003). A less obvious role of biodiversity on water provisioning is through the provision of ice nucleators (atmospheric particles that serve as condensation and ice nuclei in clouds) that lead to precipitation, a substantial fraction of it are bacteria (Christner et al. 2008). Water regulation (timing) In areas with seasonal rainfall, the distribution of streamflow throughout the year is often of greater importance than total annual water yield. This is particularly important to agricultural production (as irrigation is most important during the dry season), to hydroelectric production (as energy supply is not be possible if streamflow is insufficient), and to transportation (as waterways cannot be used if water level is too low) (Table 3). Bruijnzeel (2004) reviewed the available evidence on the effects of vegetation on the seasonal distribution of water flows, and this section draws heavily on this reference. Forest clearance often has dramatic consequences to the soil’s characteristics. Bare soil becomes exposed to intense rainfall, topsoil is often compacted by machinery or overgrazing, soil faunal activity decreases dramatically, and often part of the area is occupied by impervious surfaces such as roads and settlements (Bruijnzeel 2004, and references therein). All of these contribute to gradually reducing rainfall infiltration in cleared areas. As a result, catchment response to rainfall becomes more pronounced, with large storm runoff during the rainy season translating into lower recharging of the soil and the groundwater reserves that feed springs and maintain baseflow (deep groundwater outflow). This reduced recharge associated with deforestation typically exceeds reductions in evapotranspiration, so that overall, forest clearance leads to diminished dry season (or ‘minimum’) flows (Bruijnzeel 2004). Note that this effect is attributed to changes in the soil, and not directly to the forest loss per se. Indeed, if soil surface characteristics after clearing are maintained sufficiently to allow the continued infiltration of (most of) the rainfall, then the reduction in evapotranspiration associated with forest removal will result in an increase in dry season flow. Infiltration properties may be conserved through the establishment of a well-planned and maintained road system plus the careful extraction of timber in the case of logging operations, or by applying appropriate soil conservation measures after clearing for agricultural purposes (Bruijnzeel 2004 and references therein). Effects of deforestation on baseflows are expected to be more pronounced following severe surface disturbance in the case of deep soils with large storage capacity than in the case of more shallow soils which have little capacity to store water anyway (Bruijnzeel 2004). The same conditions that reduce water infiltration (resulting in lower baseflow) result in a higher surface runoff (saturation overland flow), resulting in higher peakflows during the wet season. However, even with minimum soil disturbance, there will still be increases in peakflows after forest removal, because the associated reduction in evapotranspiration will cause the soil to be wetter and therefore more responsive to rainfall (Bruijnzeel 2004). The fraction of water that flows as surface runoff also depends on the antecedent soil moisture 126 status, and storm characteristics: the runoff response is higher if the soils are already very wet from previous rainfall; and once the soil becomes saturated any further rainfall will increase the saturation overland flow (see section 4.13 for a discussion of the effects of forests on flooding). Reforestation may not immediately restore the modulating effects of the original forest in enhancing low flows in the dry season and reducing peak flows in the wet season, as the soil storage and infiltration capacities lost with deforestation may take years to recover. In fact, the most immediate effect of reforestation may be a decline in water yields that is particularly felt in the dry season, as a result of a rapid increase in evapotranspiration. This is particularly likely to occur for fast-growing (frequently exotic) tree plantations. Cloud forest reforestation, on the other hand, is more likely to quickly result in enhanced low flows as a result of moisture interception by the vegetation (Bruijnzeel 2004). Inland waters, such as lakes and wetlands, are traditionally considered to be very important for the temporal regulation of water flow, both by accumulating water during wet periods (reducing peak flow) and providing a reserve of water during dry periods that maintains base flow in adjacent rivers (Finlayson et al. 2005; V6rdsmarty et al. 2005). A comprehensive review by Bullock & Acreman (2003) found support for the former effect, but evidence against the latter. Evidence that wetlands have an effect in reducing or delaying floods is particularly convincing for floodplain wetlands, but less so for headwater wetlands (e.g., bogs and river margins), for which a substantial minority of studies report wetlands are associated with increased flood peaks. In contrast, most (two-thirds) of the studies concluded that wetlands reduce flow in the dry season, backed by overwhelming evidence that evaporation from wetlands is higher than from non-wetland portions of the catchment during dry periods (Bullock & Acreman 2003). Water purification (quality) Water quality is a measure of the chemicals, pathogens, nutrients, salts, and sediments in surface and groundwater. The importance of water quality to domestic use, particularly to drinking supply, is obvious (Dudley & Stolton 2003). But water quality is also very important for food production (including crops, livestock and inland and marine fisheries) and for recreational use (Table 3) (V6rdésmarty et al. 2005). Sediments reduce the storage capacity of reservoirs, thereby affecting water supply and hydroelectric production (Postel & Thompson 2005; Arthurton et al. 2007). Ecosystems with intact groundcover and root systems are considered very effective at improving water quality. Vegetation, microbes, and soils remove pollutants from overland flow and from groundwater by: physically trapping water and sediments; adhering to contaminants; reducing water speed to enhance infiltration; biochemical transformation of nutrients and contaminants; absorbing water and nutrients from the root zone; stabilising eroding banks; and by diluting contaminated water (Brauman et al. 2007). Streamside ecosystems such as riparian forests reduce nutrient movement to streams, therefore playing a key role in controlling nonpoint sources of pollution by sediments and nutrients in agricultural watersheds (Naiman & Decamps 1997). These ecosystems can trap sediments and sediment-bound pollutants in surface runoff (e.g. removing 80-90% of sediments leaving agricultural fields; Naiman & Decamps 1997) as well as absorb nutrients dissolved in the water, such as nitrogen and phosphorus (e.g. reducing local nitrate concentrations from cropland runoff by 5% to 30% per meter width; Brauman et al. 2007). 127 These buffer ecosystems can thus potentially reduce water treatment costs for downstream users. The macrophytes and microbes that promote denitrification and other biochemical processes that improve water quality are particularly abundant in wetlands, which are so reliable at removing suspended solids, phosphorus, and nitrogen from wastewater that they are regularly integrated into treatment plants (e.g. Sundaravadivel & Vigneswaran 2001). Wetland biota can also remove waterborne toxins and heavy metals from the water (e.g. Simpson et al. 1983). The effects of wetlands in environmental filtration are potentially very large. It has been estimated that converting less than 10% of the Mississippi Basin to wetlands and riparian forest would reduce 10% to 40% of the nitrogen currently creating the hypoxic zone in the Gulf of Mexico (Mitsch et al. 2001). Forests and other mature ecosystems generally improve water quality in a catchment, by reducing surface erosion and increasing water infiltration and therefore soil filtration of pollutants. Surface erosion is rarely significant in areas where the soil surface is protected against the direct impact of the rain through a litter layer maintained by some sort of vegetation (Bruijnzeel 2004). Erosion rates can increase very significantly with deforestation, as soils tend to lose organic matter and become compacted and crusty, with impaired infiltration capacity (see above). Increased surface runoff increases erosion and watershed sediment yields. Fast water flow as surface runoff also means that pollutants built up in ecosystems (e.g. through decomposition, fertilizer application) are quickly transported to rivers, rather than being filtered through the soil (Brauman et al. 2007). It seems to be the case that there is high functional redundancy in the effects of species on the provision and regulation of freshwater water, for example in the effect of plants in protecting against soil erosion. Some species or groups of species, however, may have particularly important roles. Freshwater mussels are particularly effective biofilters, filtering suspended particles (such as clay, silt, bacteria and phytoplankton and small zooplankton) and bio-deposit the particulate matter to the sediment floor, increasing water quality and clarity (McIvor 2004). Beavers are bio-engineers, creating dams that affect hydrological flow and nutrient cycling, improving downstream water quality (Naiman et al. 1986). What is the relationship between habitat area and the impact of ecosystems on the provision, regulation and purification of fresh water? It is likely that a minimum area is needed for ecosystems to have a significant effect on water provision and regulation. Studies suggest that a riparian buffer needs to be a minimum 25 m wide to remove nutrient and pollutants and a minimum of 50 m to provide detritus removal and bank stabilisation (Scherr & McNeely 2008). In an analysis of 27 US water suppliers, treatment costs for drinking water deriving from watersheds covered at least 60% by forest were half of the cost of treating water from watersheds with 30% forest cover, and one-third of the cost of treating water from watersheds with 10% forest cover (Ernst 2004; Postel & Thompson 2005). Re-plotted as a function of water treatment costs avoided (Figure 16a) this indicates somewhat diminished returns for increasing fraction of watershed covered with forest. Combining these observations with those minimum area effects, we speculate that the overall relationship between percentage of watershed covered by a particular beneficial 128 ecosystem (e.g. forests, wetlands) and the regulation of water quality is such that no benefit is noticeable for small areas, followed by a quick increase that tends to flatten out for larger ecosystem coverage (Figure 16b). a) b) Cc) no] oa 2 fo} 3 = A wo o g Ea = 8 rs) g _ (= [= € co £ = 8 8 s va s 2 $ L444 ++ + +—+ > > 20 40 60 % of watershed covered by ecosystem % of watershed covered by cloud forest % of watershed covered by forest Figure 16. Relationship between area covered by an ecosystem and its impact on the provision and regulation of freshwater. a) Relationship between percentage of watershed covered by forest and avoided water treatment costs (from Postel & Thompson 2005). b) Speculated relationship between ecosystem cover in a watershed and regulation of water timing and quality. c) Speculated relationship between cloud forest cover and water quantity. We did not find references with the shape of the relationship between regulation of water timing and ecosystem coverage, but suggest it will be similar to that for water quality, with very small or negligible effects for very low coverage and diminishing effects for high coverage (Figure 16b). As for water provision, as mentioned above the effect for both forests and wetlands seems to be one of declining total water yields for increasing ecosystem coverage. If evapotranspiration increases linearly with area, this effect might perhaps be linear. As for cloud forests, assuming that most water interception takes place in the most exposed areas of the forest (with core areas of the forest sheltered from cloud/fog contact by surrounding forest), we again speculate that the quantity of water captured is likely to increase rapidly for a small amount of forest, and then increase with diminishing returns for larger forest coverage (Figure Figure 16c; although note that we know of no studies that have looked at this). 4.9.4 What are the main threats to the provision, regulation and purification of fresh water? Water provision (quantity) Global deforestation is estimated to have reduced global vapour flows from land by 4% (3000 km*/yr), increasing water yields, although these gain s have been mainly offset by increased vapour flow caused by irrigation (2,600km*/yr) (Gordon et al. 2005). No global data seem to exist on trends in cloud forest cover in particular, but these, and their water provisioning services, are threatened by deforestation (e.g. for coffee plantations and timber extraction) and by climate change (which is raising the altitude of the cloud cap) (Bubb et al. 2004). Invasive species are in some areas compromising local water yields (e.g., Enright 2000; Zavaletta 2000). Unsustainable exploitation (mining) and pollution of water deposits, particularly of aquifers, is reducing the capacity of ecosystems to provide what should be a renewable resource (Pimentel et al. 2004). Water regulation (timing) Land conversion, including widespread loss of forests (Mayaux et al. 2005) and wetlands (Zedler & Kercher 2005) and changes in soil properties due to inappropriate soil management (Bruijnzeel 2004), are affecting the capacity of ecosystems to regulate seasonal water flow, resulting in increases in peak flows and declines in dry flows in many regions (Postel & Thompson 2005). Water purification (quality) Again, land conversion, including !oss of forests and wetlands and soil degradation, is affecting the capacity of ecosystems to filtrate and purify water. At the same time, agricultural, urban and industrial effluents are adding large quantities of pollutants aquatic systems in particular, including groundwater (Postel & Thompson 2005). Mussels are declining globally, affecting their role as biofilters, as result of widespread changes such as damming of rivers, land use change, and pollution and invasive species (Aylward et al. 2004). 4.9.5 Are abrupt changes likely in the provision and reguiation of fresh water? Most changes to the capacity of ecosystems to regulate and provide freshwater seem to derive from, and be generally proportional to, land use change. There are however situations in which a relatively small additional change may trigger a disproportionate — and sometimes difficult to reverse — response from ecosystems’ hydrological function (Gordon et al. 2008). A few examples are presented below. Human-induced eutrophication can trigger sudden shifts in lakes and reservoirs from clear to turbid due to algal blooms (Scheffer et al. 2001). These blooms may include toxic cyanobacteria and causes major problems in water treatment works, particularly those where treatment is by direct filtration (Hitzfeld et al. 2000). They also affect freshwater fisheries and recreational use of water bodies. Reduction of nutrient concentrations is often insufficient to restore the original state, with restoration requiring substantially lower nutrient levels than those at which the regime shift occurred (Scheffer et al. 2001). Eutrophication may also trigger sudden shifts in estuarine or coastal ecosystems with the creation of dead (hypoxic) zones, affecting fisheries (Rabalais et al. 2002). Deforestation on steep slopes may result in changes in soil properties that are not easily reversed as topsoil is removed by erosion, thereby making it difficult for new vegetation to get established. In particular, if gullies are not treated at an early stage, they may reach a 130 point where restoration becomes difficult and expensive, as the moderating effect of vegetation on actively eroding gullies is limited and additional mechanical measures such as check dams, retaining walls and diversion ditches become necessary (Bruijnzeel 2004). Active gully erosion substantially increases catchment sediment yields, affecting water quality as well as the storage capacity of reservoirs. Cloud forest loss may also result in a regime shift that may be largely irreversible. In some areas, such forests were established under a wetter rainfall regime, thousands of years previously. Necessary moisture is supplied through condensation of water from clouds intercepted by the canopy. If the trees are cut, this water input stops and the resulting conditions can be too dry for recovery of the forest (Folke et al. 2004). Climate change can bring potentially trigger sudden changes, particularly in regions where ecosystems are already highly water-stressed. Overall, we predict that there is a medium to high probability of localised regime shifts in the capacity of ecosystems to regulate and provide freshwater and the benefits we derive from it. Regional-scale shifts are more likely to be associated with water eutrophication. Global-scale shifts associated with climate change are perhaps" possible. 4.9.6 Can we quantify and map the global provision, regulation, and purification of freshwater, and how it might change? There does not seem to be a large-scale hydrological model that simultaneously integrates changes in water quantity, quality and timing. The Natural Capital Project (Nat Cap) is developing models that account for all three aspects of water provisioning and regulation. Nat Cap is using a tiered approach where the first models (currently under development) will use commonly available data and simple models to estimate the relative provisioning of water yield, baseflow (dry season flow) support, flood regulation, water quality and sediment yield. Higher tier (more advanced) models are being developed which will be quantitative and allow valuation. Most previous modelling efforts have focused on mapping water yield, often in order to compare with water demands for evaluating shortfalls, or the impacts of climate change on water availability. Water yield maps are usually generated from a water balance map for individual grids across the Earth’s surface, taking into account water input from precipitation, output from evapotranspiration, and water flow across cells. A number of global hydrological models exist (see D6ll et al. 2003 for a review), including: WaterGap2 developed at the Universities of Kassel and Frankfurt, Germany (Alcamo et al. 2003a,b, 2005, 2007; Dall et al. 2003); Macro-PDM, developed by Nigel Arnell (Arnell 1999, 2004); the VIC model developed at the University of Washington (Liang et al. 1994; Nijssen et al. 2001); and the WBM model developed by the Water Systems Analysis Group at the University of New Hampshire (VGrésmarty et al. 1998). All of these models are 131 driven by monthly climatic variables and they all use a spatial resolution of 0.5° (the highest resolution of available climatic input data). They vary in the way they treat variables such as soil water storage, surface runoff and groundwater recharge. They all require calibration/correction against observed water flows in order to produce a good agreement of model results, which suggests shortcomings in the models or the way they are applied. Whilst the basic hydrological concepts used are sound and have been seen to work at hillslope to river basin scale, the application at large-scales relies on gridded data that does not match well the scale of hydrological processes. While the focus of most research has been on climate change, some of these models do incorporate land use, albeit in a very crude way (e.g. categories such as agriculture, forest, scrub, wetlands). None of them seems to have been successfully validated for its capacity to simulate the effect of landuse changes on water yields. So while theoretically one can change the landuse parameters in a model and it will respond with changes in estimated river flows (e.g. Douglas et al. 2007), it is not clear how meaningful such results would be. Adequately modelling the impacts of specific changes in land use on water provision therefore requires further development. With these caveats in consideration, if it is considered acceptable to use such calibrated models with their remaining high error bands (especially in snow affected regions) then water yield is the aspects of water provision and regulation that can best be modelled globally. Global mapping of water regulation (timing) does not seem to have been attempted yet, although those models which distinguish between surface and groundwater resources may be able to be adapted to look at flood mitigation/dry season flow sustainability. The standard temporal resolution of the models is monthly, but they are typically only validated for long-term annual averages, so estimated changes in monthly monthly values (e.g. Douglas et al. 2007) are likely to be associated with high error margins. A different type of model investigates biosphere-hydrosphere interactions. These include the LPJ Dynamic Global Vegetation model, currently being developed at the Potsdam Institute for Climate Impact Research (PIK) in Germany (Gerten et al. 2004). This model estimates impacts of changes in vegetation types on the terrestrial water balance, accounting for example for the effects of changes in CO on vegetation evapotranspiration and therefore water yields. It is therefore more responsive to changes in types of vegetation cover than stand-alone hydrological models. These models have focussed on natural vegetation (Gerten et al. 2004), although an agriculture mode is now under development. Global vegetation models are less effective at modelling hydrological processes, not accounting, for example, for groundwater flows (which are key to modelling the timing of water provision). Regional models of water quality include the SWAT (Soil and Water Assessment Tool) model, developed to assist water resource managers in assessing the impact of management and climate on water supplies and non-point source pollution in watersheds and large river basins (Arnold & Fohrer 2005). The SWAT2000 includes as model components weather, hydrology, erosion/sedimentation, plant growth, nutrients, pesticides, agricultural management, stream routing, pond/reservoir routing, bacteria transport routines and urban routines. However, SWAT this relies on detailed data not available at the global scale. No global mapping/modelling of water quality seems to have been developed yet. Such models are complex in that they depend not only on water yield models but also on correctly 132 defined hydrological pathways (e.g. rapid surface and subsurface flow, versus long term groundwater). Impacts on water quality as often very localised in time and/or space, and so the models need to be able to respond to different types of pollution sources, including point (e.g. industrial) and diffuse (e.g. agricultural) sources as well as event-based (e.g. oil spill) and long-term chronic impacts (e.g., leakage from mining fields). In summary, global hydrological models are less developed for water quality and timing than for quantity, which is unfortunate as the former are precisely those aspects of water regulation that seem to be most closely linked to ecosystem extent and condition. A global valuation of the impacts of biodiversity and ecosystem change on water resources therefore requires additional modelling efforts. Ideally, the different aspects of water provision and regulation should be addressed in an integrated way (rather than by having independent models for quantity, for timing and for quality) as there are likely to be synergies and tradeoffs between these components. We therefore recommend that Phase II builds from the efforts currently being developed by the Natural Capital Project to model the effects of landuse change on water services. The hydrology module of the Natural Capital Project is based at Cranfield University, where previous work also included the development of methodologies for large scale risk assessment of diffuse source contaminants (e.g. Kannan et al. 2007a, b). h ondition and « Adequacy of scenarios The scenarios being assessed in Phase 2 need to produce: - global maps of forest cover, distinguishing cloud from other forest types - global maps of wetland distribution - global maps of agricultural production (inc irrigated) > This is our recommendation for Phase II, based on the results of this review. It does not commit the leaders of Phase II to follow it, and it does not commit the recommended research group to actually do such work. 1333) - global maps of sources of pollution, including different types of pollution sources, including point (e.g. industrial) and diffuse (e.g. agricultural) sources 4.9.7 Insights for economic valuation Assuming it is possible to predict the direction and magnitude of the impacts of land-use changes on a watershed’s hydrological services, valuation of these services needs to consider that water flow is directional, and so users are typically downstream. The specific attribute of water provision (quantity, timing or quality) and its value will depend critically on the type of use (e.g. urban water supplies, irrigated agricultural production, , etc. — Table 1). Importantly, the value of a change in total water yield, in its timing, or in quality, will depend on how far these constrain benefits — for example, the value of an extra litre of water for crop irrigation will vary depending on whether irrigation is currently practiced, and if so, whether water availability (rather than, say, soil fertility) limits crop yields. There are likely to be trade-offs within catchments, with the same change in land use (e.g. declining forest cover) increasing the provision of one hydrological service (e.g. total water yield) but reducing the provision of another (e.g. water purification). Hence, the determination of the net effect of a land-use change on the overall value of hydrological services would need to consider impacts on water quality, water timing, and water quality, and how these impacts in turn affect both downstream users and activities. To date, such comprehensive analyses have rarely been done (see Aylward 2005 for a review). 4.9.8 Some key resources e The hydrology module of the Natural Capital Project is being developed by Sue White at Cranfield University and Guillermo Mendoza at Stanford University. e The WaterGap2 (Water — Global Assessment and Prognosis) model is being developed by Joseph Alcamo at the Center for Environmental Systems Research at the University of Kassel, and Petra D6ll at the University of Frankfurt, Germany. e The WBM model is being developed by Charles Vérésmarty by the Water Systems Analysis Group at the University of New Hampshire, USA. e The Macro-PDM model developed by Nigel Arnell, currently at the Walker Institute for Climate System Research, University of Reading, UK. e The VIC model is being developed by Dennis Lettenmaier at the University of Washington (Liang et al. 1994; Nijssen et al. 2001). e The Global Water System Project (GWSP - http://www.gwsp.org/) is currently running a project to compare various global water models, including those mentioned above. e LPJ Dynamic Global Vegetation model, currently being developed at the Potsdam Institute for Climate Impact Research (PIK) in Germany. e SWAT (Soil and Water Assessment Tool), a public domain model actively supported by the USDA Agricultural Research Service at the Grassland, Soil and Water Research Laboratory in Temple, Texas, USA. 134 4.9.9 Participants Authors Hannah Peck, Ana Rodrigues, Andrew Balmford (University of Cambridge, UK) Sue White (Cranfield University, UK) Contributors Bernhard Lehner (McGill University, Canada) Sandra Postel (Global Water Policy Project, USA) Ed Tanner (University of Cambridge, UK) Acknowledgements Neil Burgess, Robert Naiman, and Charles Vérésmarty provided useful contacts and references. 135 4.10 Wild plant fibres Not developed 136 4.11 Wild medicinal plants Not developed 137 4.12 Outdoors activities related to nature Fully developed review led by an expert in the field. 4.12.1 Why is this benefit important for human wellbeing? Outdoor activities related to nature (nature-based tourism and recreation) can be categorised in numerous ways. For the purposes of this review we consider two means of categorisation; outdoor recreation vs wildlife-based tourism, and local vs. longer-distance travel. General outdoor recreational pursuits such as walking, cycling, horse-riding and the general enjoyment of being in green spaces, the countryside and wilderness areas. Beach tourism, in as much as it relies on natural landscapes and natural processes, could also be included here. These kinds of recreation rely on natural or semi- natural landscapes but biodiversity is not the focus of the attraction. They can take place almost anywhere with scenic or wilderness value, including transformed landscapes such as the English (agricultural) landscape, managed woodlands or even urban parks and gardens. Wildlife-based tourism, whether non-consumptive (e.g. photographic, birdwatching) or consumptive (hunting and fishing). This is more closely connected to biodiversity in that it has specific elements of biodiversity (i.e. selected species) as the focus of the recreational attraction. This kind of recreation usually relies on more pristine or untransformed landscapes, and is most often (but not exclusively) associated with protected areas or other designated sites or large-scale wilderness areas, including the marine environment. Local recreation, where people travel short distances to enjoy nature, is primarily likely to be general outdoor recreation and enjoyment of nature, as per the first bullet point above, although it will also include visits to nearby nature reserves by local residents. This segment of the market is likely to be more heavily affected by biodiversity loss or the loss of natural spaces, because such loss will reduce access and increase the cost of access for users. As biodiversity is lost locally there will be fewer or no alternative substitutable sites locally. Longer distance recreation, where people travel hundreds or thousands of miles to their destination, will primarily be for wildlife viewing as per the second bullet point above although beach tourism and some outdoor pursuits will also be included here. This form of recreation will be less affected by local biodiversity loss, since there will be alternative substitute sites with equivalent access or travel costs elsewhere in the world. There are many reasons why the provision of outdoor recreational activities is beneficial. Time spent experiencing nature is enjoyable, but also benefits our mental and physical health, fosters well-being and sense of fulfilment (Bird 2005; Brown & Grant 2005), and aids education (Fjortoft 2001). The psychological benefits of access to nature and natural environments, to a primarily urban, sedentary population, and the resultant improvements in productivity of the workforce, has been recognised since the industrial revolution: 138 ‘thousands of tired, nerve-shaken, over-civilised people are beginning to find out that going to the mountains is going home; that wilderness is a necessity; and that mountain parks and reservations are useful not only as fountains of timber and irrigating rivers, but as fountains of life’ (Muir 1898). There is a body of evidence that strongly suggests that contact with nature is beneficial to mental health and wellbeing (Henwood 2003), aiding patient recovery, reducing stress and anxiety (Ulrich et al. 1991), improving worker concentration and reducing crime and aggression (Bird 2007). Horticulture is widely used as a form of therapy in prisons and hospitals, and BCTV’s ‘green gym’ volunteers gain more benefits than just physical health improvements from manual conservation work (Brown & Grant 2005). Beyond the psychological are demonstrable physical benefits of outdoor recreation. Many of the outdoor activities that people enjoy rely on nature to provide the rich landscapes and stimuli that underpin these experiences, ranging from gardening and walking in the park to hiking and wildlife watching. Many of these are active pursuits, and as such are extremely valuable in tackling ill-health in the face of a modern cultural environment that encourages overconsumption of food and discourages physical activity (Hill & Peters 1998). Currently around 1.6 billion adults are overweight worldwide, with in excess of 400 million adults clinically obese, more than 115 million of whom are from developing countries (WHO). This pandemic brings with it not only serious health disorders but also social and psychological problems, all resulting in an enormous economic cost, conservatively estimated at $117 billion annually in the United States alone (Stein & Colditz 2004). However research indicates that the obesity epidemic in the United States could be halted by increasing time spent walking by only 2-3 minutes per day on average (Veerman et al. 2007). Access to green public spaces is important to urban senior citizens’ longevity (Takano et al. 2002), while accessible, large, and attractive public spaces demonstrably increase the frequency of walking, with respondents describing the importance of trees and birds amongst the features important to them (Gilles-Corti et al. 2005). Although there is currently no evidence of additional physical health benefits from exercising outdoors per se (Henwood 2002), “Health Walk’ and ‘Green Gym’ outdoor exercise schemes in the UK have demonstrated that participants are more likely to continue with this type of exercise than indoor gym-based exercise routines (Brown & Grant 2005). Finally, the tourism industry, incorporating nature-based tourism and outdoor recreation, is one of the world’s largest and fastest-growing industries, supporting millions of jobs worldwide and contributing significant proportions of the GDP of many developing and developed countries (www.worldtourism.org). Almost a century ago the scenic value of Switzerland’s natural landscapes was valued in excess of $200 million annually (Chamberlain 1910), a fact which helped to fuel the massive global expansion of national parks and protected areas and with it the birth of the modern conservation movement (Runte 1987). Nature-based tourism is one of the only non-consumptive uses of protected areas/ecosystems/biodiversity that generates tangible economic values (Ceballos-Lascurain 1996; Gossling 1999), thereby making it an important ecosystem service (MEA 2005). 4.12.2 What are the overall trends in the provision of this benefit? With increasing ease and affordability of travel, better access and communications, and increasing affluence and leisure time, the global tourism industry is growing rapidly. Travel 139 and tourism account for 10% of world GDP, 8% of jobs and 12% of global investment annually (WTTC 2007). Tourism also has the highest potential for growth of any industry — currently running at more than 4% per annum, with an average of 3% in developed economies and more than 7% in emerging markets” (WTTC 2007). Within this, nature- based tourism and recreation is considered to be an increasingly significant sub-sector, although there are little consistent aggregate data (Goodwin 1996; Eagles 2002). Although there have been numerous studies of the magnitude (Goodwin et al. 1998; Eagles et al. 2000; Wade et al. 2001) and value (e.g. Wells 1993; Moran 1994; Shrestha et al. 2007) of nature-based tourism at local and national scales, to date there has been no attempt at a global assessment. This is in part because there is no consistent approach to recording and reporting visitor arrivals to parks and protected areas (Eagles 2002) and other forms of outdoor recreation go largely unrecorded in any form other than periodic site-based or regional surveys. Although recent high-profile research suggests that nature recreation is declining per capita in US and Japan (Pergams and Zaradic 2008), this trend is not mirrored in much of the rest of the world where growth in visitation to protected areas is growing at least as fast as international tourism as a whole (Balmford et al., unpublished data). Less data are available for other types of outdoor activity, though it has been estimated that each year over half the population of the UK makes over 2.5 billion visits to urban green spaces (Wooley & Rose 2004), and 87 million Americans participated in wildlife-related recreation in 2006, an increase of 13% over the decade (USF WS 2007). As people show a universal preference for natural environments over built environments, an increasing amount of outdoor recreation would be expected with projected increases in world population and growing urbanisation (de Groot et al. 2005). This will place an increasing reliance (and therefore value) on remaining natural landscapes. Two other trends are worth noting. The first is a shifting emphasis in global tourism towards emerging, developing world destinations, and with it an increase in domestic tourism in the developing world as urban incomes rise. The second is an increasing diversification of the agricultural sector in Europe and beyond into recreational service provision. At the same time community-based tourism in rural areas in the developing world is expanding nature-based tourism beyond protected area boundaries (Walpole & Thouless, 2005). Whether these trends are increasing the overall magnitude of nature-based recreation is unclear, but they are undeniably expanding the area and range of ecosystems which provide recreationa! opportunities. 4.12.3 How is the provision of this benefit/process affected by changes in wild nature? There are almost no reliable quantitative data on the extent to which changes in biodiversity influence recreational benefits. The value derived from nature by those experiencing and using it for recreational purposes is influenced by a wide range of factors, of which biodiversity is one. Whilst it underlies the provision of key elements of a recreational experience, and of supporting services, its marginal value is difficult to determine and varies depending on the type of recreation and the perspective of the individual user. The following sub-sections review some areas where biodiversity is likely to play a role. 140 Provisioning of beautiful landscapes/seascapes Outdoor recreation users rely on the provisioning of aesthetically pleasing landscapes as a stage for their experience (De Groot et al. 2005), landscapes which are created and regulated as part of naturally functioning ecosystems. Examples include the deposition of the calcium carbonate tests of foraminiferans and reef-building organisms which, together with mollusc shells, supply material for many beaches (Yamano 2000). Studies have shown a clear preference for natural landscapes over urban ones, and people tend to prefer greener, healthy ecosystems, but more detailed judgements of landscape quality such as ‘wildness’ are likely to be more culture-dependent (de Groot et al. 2005). They are also activity-dependent. Much outdoor recreation is physical in nature and relies on the physical (slope, terrain) rather than biological (species and habitats) qualities of the landscape. Moreover landscape ‘beauty’ is in the eye of the beholder and for some can be enhanced by modification (the English countryside being a good example) which may reduce biodiversity or reduce the importance of biodiversity in the recreational experience. A study of forest recreational user preferences in the UK revealed the importance of man- made facilities for enhancing use and enjoyment amongst general and activity-based forest visitors, with only dedicated nature watchers prioritising access to biodiversity (Christie et al. 2007). The creation of the natural landscapes that make outdoor pursuits attractive is clearly a key benefit provided by biodiversity, yet there is currently insufficient knowledge to be able to define which attributes of the landscape are key to peoples’ experiences, let alone to quantify links between specific aspects of biodiversity and outdoor benefits. Typically such benefits are linked only to very general descriptions of nature such as visual complexity or perceived naturalness (Brown & Grant 2005), which are difficult to translate into biological terms or a concrete metric of biodiversity. However, land use change from natural to intensively cultivated (or urbanised) landscapes will clearly reduce or eliminate the recreational amenity value of an area. This will impact on the provision of local outdoor recreational opportunities where access/distance is a constraining factor. Provisioning of focal attractions: species diversity or presence of particular species Humans derive great enjoyment and benefit from interaction with and observation of wild animals and plants, clearly demonstrated in our love of gardening, and the popularity of wildlife watching. E.O. Wilson has argued that the desire for such experiences is the product of an innate love of living things which he terms “biophilia’, which drives our need to seek connections with other life-forms (Wilson 1984). Charismatic species such as gorillas and cetaceans clearly impact demand for outdoor recreation, even generating tourism industries in areas not previously seen as a destination, such as Rwanda. The annual recreational value of wildlife watching in Lake Nakuru National Park, Kenya has been estimated at between US$7-15 million, with flamingos creating more than one third of the value (Navrud & Mungatana 1994). The reintroduction of wolves to Yellowstone has attracted additional tourists to the national park, with consequent economic and social benefits estimated at between US$6-9 million per year (Donlan et al. 2006). Interestingly, the presence of the wolves alone increases the enjoyment of the park experience for tourists, regardless of whether or not they see any wolves (Montag et al. 2005). 141 Charismatic bird species have a significant influence in generating wildlife tourism in the UK, with an estimated 290,000 people visiting osprey nesting sites each year, bringing an extra £3.5 million to local economies, while the presence of sea eagles on Mull attracts an extra £1.4-1.6 million of spending each year (Dickie et al. 2006). In a study of nature tourists visiting Uganda Naidoo & Adamowicz (2005) found a tangible demand for bird diversity and a willingness to pay for it. Similarly, in Komodo National Park in Indonesia, visitors’ willingness to pay increased entrance fees was positively correlated with the number of dragons they had seen during their visit (Walpole, 1997). Biodiversity is perhaps even more significant in marine tourism. There are around 10 million active scuba divers who generate around $1.2 billion annually, and reef-related tourism is growing rapidly at around 20% per year (Cesar et al. 2004). For these outdoor users the diversity of species found on a coral reef is clearly paramount to the enjoyment of a diving experience. In the aftermath of the mass coral bleaching event of 1997-8 this was demonstrated by the losses to the dive industry in regions hit by coral bleaching and the associated loss of reef species, for example in Palau, where as much as 10 percent of the diving industry’s producer surplus was lost (Schuttenberg 2001). A survey of tourists in Bonaire revealed that 80% would not be willing to return in the event of coral bleaching (Uyarra et al. 2005). These examples all reveal the importance of biodiversity, in terms of species presence (which often relies on broader ecosystem health). It is clear that wildlife-based recreation is generally more sensitive to the availability of pristine environments, species abundance and diversity than broader outdoor recreation. Moreover some of the highest profile attractions, the top predators and other large animals, require intact, functioning ecosystems to persist. Yet the relationship between changes in these elements of biodiversity and the provision (and value) of recreational benefits is complex, case specific, rarely linear and sometimes counter-intuitive. Special features of tourism as an ecosystem service There is no linear relationship between magnitude of visitation, expenditure or other measure of value and biodiversity. A reduction in bird diversity in a tropical forest may impact on bird-watchers who are motivated by, amongst other things, species richness. Likewise a decreasing chance of seeing a rare species such as the Komodo dragon, or the disappearance of any of the ‘big five’ African mammal attractions might deter some from visiting a particular park. However in general there is likely to be significant elasticity of demand, with a change in number of flamingos, or wildebeest, or lions in a park having little immediate impact on visitor numbers (Walpole & Thouless 2005). In other cases increasing rarity and risk of extinction may increase tourism demand for certain sites and species, with the mountain gorilla, tiger and giant panda as obvious examples (Entwistle et al. 2000). Tourism is still growing and is not currently limited by availability of natural attractions. Increasing availability of nature-based tourism attractions (for example due to the establishment of new protected area destinations} increases choice, substitutability and competition between destinations. It may thus alter the geographic distribution of economic benefits from tourism as new destinations capture a portion of the market. However it will not necessarily increase the overall value of tourism at global scales (Walpole & Thouless, 2005). 142 The value of tourism as a rationale for ecosystem conservation is limited by the size of the (albeit growing) global market, but in the short term is unlikely to be constrained by a general reduction in biodiversity and availability of recreational sites until such point as access, overcrowding or species disappearance cause a tipping point to be reached. Such tipping points may be reached at the local level, particularly in terms of access to general outdoor recreation where travel cost is likely to be a limiting factor. In theory, if local sites for outdoor recreation decline (with increasing urbanisation or agricultural expansion), there will be less opportunity for people to obtain this benefit and welfare will decline. Overall welfare will not decline as long as substitutable opportunities are available elsewhere. This is generally more likely to be the case for longer-distance wildlife tourism than for more general outdoor recreation. Most tourists will simply choose an alternative destination. Unlike other ecosystem services, tourism benefits do not flow to the beneficiary but rather the reverse — beneficiaries travel to the point where the service is produced. For international tourism at least, there is not the same distance or access constraint as for some provisioning, regulating or cultural services, and substitutability is possible. Provisioning of regulatory services Tourism and recreation rely on a range of regulatory and supporting services provided in part by biodiversity, including climate regulation, water flow, waste management, coastal protection and erosion control. Without the many regulating services provided by wild ecosystems attractive environments for outdoor recreation would not be maintained. This has been seen in some coastal areas in Mauritius, where seagrass beds were removed in the belief that they were not aesthetically pleasing to tourists, resulting in a worse environment for swimmers, with increased turbidity and loss of infaunal biodiversity (Daby 2003). The role of biodiversity in these ancillary services are dealt with elsewhere in this review and thus are not considered further here. 4.12.4 What are the main threats the provision of this benefit? With increasing global population, urbanisation and land transformation the provision of remaining natural landscapes will increase in importance, particularly for the growing proportion of humanity that does not have any exposure to natural settings. Many natural environments, such as coral reefs, wetlands and montane environments are already becoming significantly degraded and transformed. Coastal areas are also under threat from erosion, over-development and sea level rise. Intensifying conversion of wild habitat to agricultural land will drastically impact outdoor recreation for many, particularly where access to outdoor recreation becomes constrained, although there is insufficient knowledge of landscape preferences to predict the net effects on all of humanity. At the same time the proportion of the world’s surface under designated protection continues to increase (UNEP-WCMC, 2008), and natural areas continue to be added to the UNESCO World Heritage List (http://whe.unesco.org). Although such designation does not guarantee protection (and many protected areas and wildlife populations remain threatened by development pressures, poaching, logging, settlement, etc) it does raise profile and makes such areas more likely to become tourism destinations. This is in effect increasing the provision of relatively pristine natural environments upon which much nature-based tourism is based. 143 Globally 32% of amphibian species, 23% of mammals, and 12% of bird species are threatened with extinction, while the status of reptiles, plants and fish is uncertain. Within these groups, charismatic taxa such as primates, carnivores and albatrosses are particularly at risk. In recent years, the extinction risk for birds and amphibians has increased (Baillie et al. 2004) and the pattern is likely to be mirrored in other taxonomic groups. WWF’s Living Planet Index also suggests that, overall, wild species abundance is also in decline, implying that even the less threatened wildlife may become harder to find. 4.12.5 Are abrupt changes likely in the provision of this benefit? Abrupt changes in the provision of tourism benefits can occur for a range of reasons. Some of these may be ecological, as systems reach tipping points. Key wildlife populations may collapse through disease or other factors, fire may destroy picturesque landscapes, corals may bleach with sudden temperature shifts, systems may suddenly change from one (attractive) to another (less desirable) stable state. Some of these will be reversible, others may be more permanent. Abrupt shifts may also (and perhaps more often) be socially instigated. War, terrorism, socio-political disruption, natural disasters and health crises all tend to rapidly and negatively affect international tourism demand, as evinced for example by Kenya, Zimbabwe, Bali, Egypt and Nepal in recent times. Likewise events such as the foot and mouth outbreak in the UK in 2001 may have dramatic impacts as people are prevented from visiting the countryside for recreation. In some cases it has also been shown that abrupt changes in management of areas, such as degazetting parks or increasing entrance fees, can result in a sudden drop in recreational demand (Goodwin et al. 1998). The current rapid rise in oil prices (and thus aviation fuel costs) and potential carbon taxes on flying, may have similar impacts on international tourism if such changes are too sudden. 4.12.6 Can we quantify and map the global provision of this benefit and how it might change? We are a long way from being able to map the production of benefits from nature-related outdoor recreation based on changing states of biodiversity. Firstly there is little known about the link between features of biodiversity and outdoor recreation demand on which to base such models. Indeed the analysis of outdoor recreation in general is hampered by a great scarcity of data on outdoor users, and where monitoring does exist is often unreliable and patchy. Secondly and in such cases where the link has been demonstrated or recreational value estimated, it is not clear how far we can generalise such a relationship from very localised case studies. In contrast to the general terms such as ‘perceived naturalness’ which are used to describe the desired aspects of landscapes for general outdoor recreation in some literature, valuation studies focussing on the importance of particular species or ecosystems to certain types of outdoor visitor are too context specific to allow generalisation. It is difficult to tie down universal preferences for biological attributes from such a diverse group as nature-based recreational users, so linking biophysical attributes to amenity value globally is problematic. Part of the difficulty may lie in that such preferences are context- specific, that is, they are dependent on the physical and cultural environment from which 144 people originate. It has been found that tourists’ perception of environmental quality is influenced by their past experiences, socio-economic background and culture (Petrosillo et al. 2006). Thirdly, most studies focus on total economic valuation rather than marginal values and so the application of scenarios to model the impacts of change is constrained. Already developed integrated models Existing models of nature tourism demand are few and far between, and are unable to include any aspects of biodiversity amongst their variables. Brainard and colleagues (2001) have produced models predicting demand for recreation in English woodlands based on accessibility, local population, and facilities present. Likewise Jalale (1993) used conventional demand modelling to identify factors affecting visitation to national parks in Malawi and found access and facilities to be important. Balmford and colleagues (unpublished) have been the first to attempt to model visitor numbers to protected areas globally, using a database of site-specific and national data. None of these models were able to include significant variables relating to biodiversity. This may reflect both the importance of non-biotic factors (such as access), as well as the range of motivations for nature-based tourism, including various aspects of naturalness and biodiversity that cannot easily be distilled into adequate quantitative variables (Goodwin & Leader-Williams 2000). In the UK Carver and colleagues (2002) have developed a method for mapping a subjective judgement of landscape, in this case wilderness. By allowing users to weight the importance of underlying spatial variables such as distance from nearest road, the different preferences of various groups for landscape qualities could be determined, and potentially the relative importance of those qualities produced by biodiversity could be measured. It is possible that a global map of a general subjective descriptor such as ‘natural attractiveness’ could be pieced together in this manner by allowing local weighting of the underlying landscape qualities. Such a map could then be combined with other recognised variables influencing tourism demand such as accessibility to create a model for general domestic outdoor recreation potential. For wildlife-based tourism it will probably be necessary to create individual models for clearly distinct categories such as scuba divers and bird watchers, and to limit any model extrapolation to those areas where tourism currently exists (e.g within national parks rather than across natural ecosystems more broadly. The challenge is that likely (and relatively simplified) scenarios based on land use change outside protected areas would not trigger marginal changes in protected area tourism that could be valued, unless models included the impact of neighbouring land use change on protected area biodiversity. Even if it did, the likelihood of substitutable alternative destinations as menetioned above would limit the impact on global provision of tourism benefits. Availability of adequate data e Further data on the magnitude (and value) of different types of tourism in different ecosystems is a pre-requisite for any modelling exercise. e Further work examining how nature-based recreation demand (either general outdoor recreation with a primarily local catchment or nature-based tourism with a potentially much broader regional or global catchment) responds to changes in 145 biodiversity (or land use type at the very least) is required before any global scenario models can be created. Adequacy of scenarios e None currently available for nature-based tourism and recreational activities. What can be done in Phase II? At what cost? By whom®? We recommend that in the short term, activities are two-fold, both focusing on the value of specific recreational types in different habitats. First, more in-depth literature-based syntheses of well-studied activity subtypes, namely scuba-diving and bird watching, where the relationship between recreational benefits and biodiversity is more clearly defined and where valuation studies are more advanced. Second, and of potentially greater value, modelling change in the availability of local outdoor recreation with changing land use/urbanisation and thus increasing distance to recreational sites. The importance and impact of access/distance on tourism demand is already well modelled, and numerous organisations in the UK have expertise in this area. There is also benefit in modelling tourism demand and marginal value for terrestrial protected areas. This work has recently begun in Cambridge, and investment will be required in two areas: (1) refining the global visitor model for protected areas, and (2) extending it to include economic valuation. This will require improving the coverage and comparability of protected area visitor and finance data, besides a more targeted literature- search for valuation case studies. For this, UNEP-WCMC in partnership with Cambridge University, the International Centre for Responsible Tourism and the IUCN WCPA are well-placed to lead. These two approaches would need a minimum of 24 researcher-months, and could in theory and in combination provide some kind of partial marginal valuation that could be included in a global scenario modelling exercise. 4.12.7 Insights for economic valuation The valuation of recreation can be done based on actual financial flows, although distinguishing what proportion of expenditure during a tourism excursion should be attributed to natural attractions can be challenging. Where there is no direct expenditure, other evaluation methods such as travel cost can be used. Since a propotion of the value of recreation is not captured in markets it is common in recreational valuation studies to use contingent valuation methods to estimate ‘user © This is our recommendation for Phase II, based on the results of this review. It does not commit the leaders of Phase II to follow it, and it does not commit the recommended research group to actually do such work. 146 surplus’, the proportion of the value of the experience retained by the tourist rather than captured in entrance fees, for example. Such methods are problematic, but the issue of where to draw the line in defining what to include in an evaluation of economic benefits is a very important one to resolve at the outset of any global exercise. 4.12.8 Some key resources Web resources are extremely limited since, apart from UNEP-WCMC and the University of Cambridge (and the Natural Capital Project) no one is exploring this issue at a global scale. The following is broken down into some key organisations and some web-available literature: Organisations e Natural Capital Project: www.naturalcapitalproject.org/ e World Database on Protected Areas: www.unep-wemc.org/wdpa/ e World Commission on Protected Areas: www.iucn.org/wcpa e International Centre for Responsible Tourism: www.icrtourism.org/ e World Travel Organisation: www.world-tourism.org/ Reports, publications and journal articles e RSPB. 2007. Wellbeing through Wildlife in the EU. http://www.birdlife.org/eu/pdfs/Wellbeing EU final version 2mb.pdf e Pretty J, Griffin M, Peacock J, Hine R, Sellens M and South N. 2005. A countryside for health and wellbeing: The physical and mental health benefits of green exercise. University of Essex, Colchester. Available online at: http://www.countrysiderecreation.org.uk/pdf/CRN%20exec%20summary.pdf http://195.92.230.85/Images/Hine_tem2-30031.pdf e Bird W. 2004. Natural Fit: Can Green Space and Biodiversity Increase Levels of Physical Activity? RSPB. The RSPB. Available at: http://www.rspb.org.uk/Images/natural_fit_full version _tem9-133055.pdf e Collins, S. 2006. The Makuleke model for good governance and fair benefit sharing Steve Collins in IUCN. Policy Matters http://www.iucn.org/themes/ceesp/Publications/newsletter/PM14- Section%201V.pdf e Health Council of the Netherlands. 2004. The influence of nature on social, physical and psychological wellbeing. Part 1: review of current knowledge. Report to the Minister of Agriculture, Nature and Food. http://www.rmno.nl/files_content/Nature%20and%20Health.pdf e UNEP/CMS Secretariat (2006). Wildlife watching and tourism. Bonn. Available at: http://www.cms.int/publications/pdf/wildlifewatching_text.pdf 4.12.9 Participants in this sub-review Authors James Beresford (University of Cambridge, UK) Matt Walpole (UNEP-WCMC, UK) 147 Contributors Andrew Balmford, Ana Rodrigues (University of Cambridge, UK) 148 4.13 Regulation of natural hazards Not developed 149 4.14 One-off use benefits Not developed 4.15 Non-use benefits Not developed 151 Global climate regulation This section is a very brief overview based on a quick literature review. It did not receive contributions from, nor has it been reviewed by, experts in this field. 4.15.1 Why is global climate regulation important for human wellbeing? There has been extensive work on the ways in which climate change influences human wellbeing (and therefore the ways in which a reduction of the effects of climate change affects wellbeing). It is very well-established that climate change will have profound effects on wellbeing including through effects on food and water provision, health, property and infrastructure, global security and impacts on ecosystems and the benefits we derive from them. A disproportionate burden of these impacts falls on poor regions of the world, and so climate change is predicted to affect global development goals very significantly. Key references include: e Stern (2007) — Stern Review on the Economics of Climate Change. This review has collated information on how climate change affects: water provision, food provision (including marine fisheries), health, land (including effects of sea-level rise), property/infrastructure, and the environment (including species extinctions). It further analysed how climate change affects development at global and regional scales. e House et al. (2005) — the section of the Millennium Ecosystem Assessment: Current State and Trends on Climate and Air Quality, particularly the sub-section 13.6 on Impacts of Changes in Climate and Air Quality on Human Well-being. e [PCCWGII (2007) — Climate Change 2007 — Impacts, Adaptation and Vulnerability (Working Group II contribution to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change), including an assessment impacts on fresh water resources, ecosystems, food, coastal systems, human health, industry and an overview of impacts per continent. 4.15.2 How is global climate regulation affected by changes in wild nature? There has also been extensive work on the mechanisms driving climate change, including the ways in which natural ecosystems affect the concentration of greenhouse gases, and the ways in which human activities are impacting those processes. It is well-established that terrestrial and aquatic ecosystems play key roles in the global geochemical cycles affecting the concentration of greenhouse gases in the atmosphere. Ecosystems also influence climate at a regional scale by affecting radiation (e.g. albedo), cloudiness, and surface temperature. Accordingly, it is also well-established that anthropogenic changes on ecosystems have significant impacts on climate change. Land use changes affecting forest ecosystems (deforestation, reforestation, afforestation) are particularly influential to the balance of carbon emissions, sequestration, and storage. Forest fires also contribute to climate change by releasing aerosols. Other ecosystems also play important roles including: wetlands and reservoirs (with net methane emissions); grasslands (with variable degrees of underground carbon storage); peatlands and mires (important carbon storages); agricultural lands (particularly nitrous oxide emissions, also methane). Oceans are substantial carbon 152 reservoirs and sinks, but these seem to be stable (presumably not currently being affected by human activities?). There has been plenty of work on the quantification of carbon storage, sinks and emissions for different ecosystems (including above and below ground) and these can be mapped at the global scale. Key references include: e SCBD (2003) report Interlinkages between biological diversity and climate change: Advice on the integration of biodiversity considerations into the implementation of the United Nations Framework Convention on Climate Change and its Kyoto protocol discusses ways in which climate change affects biodiversity and also ways in which changes in natural systems affect greenhouse gases emissions. e IPCCWGI (2007) — Climate Change 2007 — The Physical Science Basis (Working Group I Contribution to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change), in particular chapter 7 on Couplings Between Changes in the Climate System and Biogeochemistry. e IPCCWGIII (2007) — Climate Change 2007 — Mitigation of Climate Change (Working Group III Contribution to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change) in particular chapter 9 on Forestry. e Foley et al. (2005) includes a discussion of the ways in which land use change affects global climate. e IPCC (2001), the IPCC Special Report on Emissions Scenarios, particularly Chapter 3: Scenario Driving Forces which discusses agriculture and land-use emissions and how these affect climate change scenarios. e Pregitzer & Euskirchen (2004) discuss the role of forests in carbon cycling and storage. e McGuire, A. D., Sitch, S., Clein, J. S., Dargaville, R., Esser, G., et al. (2001) Carbon balance of the terrestrial biopshere in the twentieth century: Analyses of CO2, climate and land use effects with four process-based ecosystem models. Global Biogeochem Cycles 15, 183-206. e 3). Tian, H., Melillo, J. M., Kicklighter, D. W., McGuire, A. D. & Helfrich, J. (1999) The sensitivity of terrestrial carbon storage to historical climate variability and atmospheric CO2 in the United States. Tellus Ser B Chem Phys Meteorol 51, 414-452. 4.15.3 Can we quantify and map the value of wild nature for global climate regulation, and how it might change? There has been important work on the calculation of the economic implications of climate change and the costs of climate change mitigation. This information is fundamental for deriving a price for carbon, which can be created through both tax and trading. The first international emissions trading scheme was created by the European Union in 2006. A carbon price can be used to quantify the economic consequences of different policy actions 153 by assessing how they differ in terms of carbon storage, emission and sequestration by different ecosystems. Key references include: e Stern (2007) — Stern Review on the Economics of Climate Change, particularly section 15: Carbon Pricing and Emissions Markets in Place. 4.15.4 Participants Authors Hannah Peck, Ana Rodrigues, Andrew Balmford (University of Cambridge, UK). 154 4.16 Unknown benefits or processes Not developed 155 4.17 Scenarios Scenarios provide a methodology for exploring alternative future environments in which today’s decisions might be played out. In practice, scenarios are stories, written or spoken, built around carefully constructed plots often termed narrative storylines. Scenarios are not predictions; instead, scenarios are an approach to help manage the inherent uncertainties of decisions by examining plausible, internally consistent alternatives of how the future might unfold and comparing the potential consequences of decisions in different future contexts (SEI, n.d.; Verbug et al. 2006). Scenarios may be purely qualitative, or may generate quantitative results through the use of one or more simulation models. The generation of appropriate scenarios is a key component of Phase 2. Without these, the relevance and accuracy of all results obtained will be questionable. Throughout we refer to two scenarios, but many more can and should be generated to assess different policy packages. It is, however, fundamental that in any quantification of the economic consequences of biodiversity loss and ecosystem degradation, two scenarios are compared. These need to be identical in everything else but the specific policy package being tested. Otherwise, the economic results cannot be directly attributed to a difference in biodiversity/ecosystems. For example, an appropriate contrast would be between the state of the world by 2025 generated by a business-as-usual (BAU) scenario, and an alternative state of the world generated by an otherwise identical scenario that includes the policy option of “a comprehensive global network of marine protected areas”. These two scenarios could be contrasted to evaluate the specific economic consequences (both the costs and the benefits) of implementing the marine protected area network, as everything else (e.g.. population, technology, consumption) would be the same. It would not be appropriate to try to evaluate the consequences of establishing a comprehensive marine reserve network by contrasting a 2008 state of the world (with its incipient network of marine protected areas) with a 2025 state of the world generated by the policy option scenario. Such a contrast would not take into account the changes in many other variables, including total human population, demand for marine fish, technological changes in fishing methods, etc. In order to be useful, the scenarios need to have the right level of information at the right spatial scale. For example, if the action is to implement new protected areas, the scenario needs to state where they are; if the action is to manage fisheries appropriately, the scenario needs to spell out how that would be done, for example by creating no-fishing zones and setting sustainable fisheries quotas. If the scenarios do not have the necessary detail, they will be inappropriate to test how changes in biodiversity/ecosystems affect the benefits we derive from them. Scenario development has become a popular tool for the assessment of policy options, at scales ranging from national and regional (e.g. Ferreras et al. 2001; Green et al. 2005; Johansson et al. 2007; Rounsevell et al. 2006; Sandker et al. 2007; Soares-Filho et al. 2006) to global (e.g. House et al. 2002; Kindermann et al. 2006; Msangi et al. 2007; Sathaye et al. 2003; Ten Brink et al. 2007) and there is a growing literature on scenario generation and implementation. Recent large-scale scenario exercises likely to be relevant to the present assessment include: ATEAM (for Europe); the global and regional scale the Millennium Ecosystem Assessment; Global Environment Outlook exercises; the International Assessment of Agricultural Knowledge, Science and Technology for Development; the 156 Intergovernmental Panel on Climate Change fourth assessment; World Energy Outlook; and the OECD Outlook series (environment, agriculture, economics). It is beyond the capacity of the current project to review the extensive literature on scenario development, to recommend which types of methods/approaches for generating scenarios are the most suitable for the current assessment, or to propose in any detail which specific scenarios should be contrasted. We emphatically recommend that this is a major priority for Phase 2 of the Review, the task that underlies the development of all other work. For each review theme, we have made a general assessment of whether the scenario results currently being generated in global to regional scale modelling exercises are likely to produce the adequate level of information needed for the biodiversity response models/approaches that we have proposed for quantifying changes in benefit/services provision. In addition, we provide a brief overview (not an exhaustive list) of main scenarios that we hope covers the major models used in global scale scenario exercises. Table 4 Main scenarios used in policy evaluation Model / approach Global climate models — Hadley Centre, CSIRO, etc. Example scenario exercises* IPCC (Nakicenovic & Swart 2000) Output variables Climate variables such as temperature, precipitation, relative humidity Spatial scale of future simulation Multiple degrees - downscaling approaches permit finer scale results Theme relevance Scope to improve Ongoing scope to improve integrated land cover models Reference Parry et al. 2007; Solomon et al. 2007 GEO; regional exercises Land use including biofuel production; agricultural production Surface runoff and river discharges Air quality — pollutant emissions (NOx, SO2, CO) Land use demand, land cover 5 degree grid — Asia- Pacific region 5 degree grid — Asia- Pacific regi Grid (SGale 8) — Asia- Pacific region Flexible — depends on inputs 157 (Land use / cover) (Freshwater) (Land use / cover) Finer resolution NIES 2008 Verburg e¢ al. 2002 EcoSim / EcoPath GLOBIO GUMBO IMAGE GEO, GBO, OECD Environment Outlook ‘Big Gov / Eco-topia / Mad Max / Star Trek’ set — Costanza services (see simulation of (2000) model services and Carbon price scenarios GEO, GBO, IPCC, MA, IAASTD variables: other models | resolution CO2 emissions Fish production; biomass Freshwater fisheries simulation Pauly et al. 2002 Biodiversity Additional Alkemade et indicator indicators al. 2006 “mean may be abundance of developed original species’ Change in 0.5 degree (Land use / Modelling ecosystem (fractional cover) approach is extent cover per relatively cell) basic Land use Global (11 (Land use / biomes) cover) Boumans et al. 2002 Add spatial component Ecosystem Direct description their below) marginal value National scale Multiple socio- economic variables (economics, food demand, population, etc) Useful to More direct Hughes provide environmental | 2001; internally limits Hughes & consistent Irfan 2007 sets of drivers for other models Environment Various; but | Finer remaining typically fossil fuels, have finer area of forested resolution land, water usage, and Kindermann et al. 2006 Forest area & (Land use / Finer biomass cover); resolution carbon Finer Bouwman et resolution al. 2006 (Land use / cover) Land cover — 0.5 degree including cropland, pastureland, forest, urban... Soil 0.5 degree (Land use / degradation cover) NPP 0.5 degree Carbon storage 158 Timber 24 regions Forest production of world (?) | services Nitrogen 0.5 degree deposition Risk of water 24 regions (Land use / Potting & erosion of soils |} (?) cover) Bakkes 2004 IMPACT Crop, livestock | 69 regions (Land use / Finer Rosegrant et consumption of world cover) resolution al. 1995 and production Fish Fisheries consumption and production River runoff 0.5 degree (Freshwater / | Integrate Arnell 2003 marine) effects of land cover change Exports of River basin | Fisheries; Linkage with | Seitzinger et nutrients from mouths oceanic oceanic al. 2005 rivers to carbon transport coastal zones sequestration | model WaterGAP Water flow and | 0.5 degree Freshwater D6ll et al. storage 2003 (surface runoff, groundwater recharge, river discharge, water storage in soil, groundwater and surface water bodies) Water use 0.5 degree Freshwater (withdrawals for irrigation, livestock, households, thermal power plant and manufacturing) * Acronyms — scenario exercises GEO — Global Environment Outlook (4) GBO — Global Biodiversity Outlook (2) 159 IAASTD - International Assessment of Agricultural Knowledge, Science and Technology for Development IPCC - Intergovernmental Panel on Climate Change assessment reports (4°) MA — Millennium Ecosystem Assessment — Carpenter et al. 2005 Model descriptions (Quotation marks indicate text directly from paper or webpage describing model) Global climate models. Atmosphere-Ocean General Circulation Models designed to reproduce current and historical climate, and project the response of future climate to changes in variables such as the Earth’s atmospheric composition, land cover and solar forcing. Confidence is higher for changes in temperature than in precipitation. See IPCC reports for further details. AIM -— Asia-Pacific Integrated Model An integrated assessment model, designed to assess policy options for stabilising and adapting to changes in global climate, particularly in the Asian-Pacific region. Developed at National Institute for Environmental Studies, Japan. CLUE /CLUE-S Developed for the analysis of land use at a national to continental scale. CLUE-S is a variant for smaller regional scale applications (e.g., a watershed or province) at a fine spatial resolution. Demand is simulated with one module, and spatial distribution with another — it is possible to supply this second module with demands from a different external scenario. Model uses decision rules such as conversion probabilities between land uses. Developed at Wageningen University. EcoSim/ EcoPath / EcoSpace Simulates biomass within marine food web models under the impacts of competing fishing fleets. EcoSpace simulates the influence of protected area networks. Developed at University of British Columbia. GLOBIO3 Developed for the analysis of impact of policy options on biodiversity. Indicator is “mean abundance of original species of ecosystem’ — hence a relative indicator of biodiversity compared to a reference ‘natural’ state, rather than an absolute indicator of biodiversity compared between locations. Based on a set of cause-effect relationships grounded in the literature on the impacts of pressures on biodiversity. Developed at the Netherlands Environment Assessment Agency / UNEP/GRID Arendal / UNEP-WCMC. 160 Gumbo - Global Unified Metamodel of the Biosphere “Gumbo consists of five sectors or spheres: Atmosphere, Lithosphere, Hydrosphere, Biosphere, and Antrophosphere (human systems). It is also divided into 11 biomes or ecosystem types which encompass the entire surface area of the planet: Open Ocean, Coastal Ocean, Forests, Grasslands, Wetlands, Lakes/Rivers, Deserts, Tundra, Ice/rock, Croplands, and Urban. Global energy, carbon, nitrogen, and water dynamics are incorporated in Gumbo, along with soil erosion and formation, fossil fuel formation and use, plant and animal productivity and harvest, human population, energy, biomass, and water use, economic production and welfare, and changes in natural, built, human, and social capital”. Simulates ecosystem services on a global scale, based on flows and storages in the model. The services are: gas regulation, climate regulation, disturbance regulation, water use, soil formation, nutrient cycling, waste treatment, food production, raw materials, and recreation/cultural. Developed at University of Vermont. IFs - International Futures “The broad purpose of IFs is to serve as a thinking tool for the analysis of near through long-term country-specific, regional, and global futures across multiple, interacting issue areas”. The overall model incorporates different sub-models, including: Population, Economics, Agriculture, Education, Energy, Socio-Political, the International Political, Environment, Technology, and Health. Model relies on simulation of macroagents (e.g. governments) and markets. Developed at University of Denver. IIASA forest model [precise name unclear] “Calculates differences in net present value of different land uses using a spatially explicit integrated biophysical and socio-economic land use model. Key model parameters, such as agricultural land use and production, population growth, deforestation and forest product consumption rates were calibrated against historical rates. Land use changes are simulated in the model as a decision based on a difference between net present value of income from production on agricultural land versus net present value of income from forest products. Assuming fixed technology, the model calculates for each 0.5° grid cell the net present value difference between agricultural and forest land-uses in one-year time steps” IMAGE - Integrated Model to Assess the Global Environment The IMAGE model simulates direct and indirect pressures on human and natural systems resulting from industry, housing, transport, agriculture and forestry. Socio-economic activities and drivers of change are allocated to 24 regions of the world, and the climate, land-cover and land-use change-related processes are represented on a 0.5 degree grid. Developed at the Netherlands Environment Assessment Agency. IMPACT - International Model for Policy Analysis of Agricultural Commodities and Trade IMPACT has as its basis a set of national supply and demand equations for 32 commodities, with each country model being linked to the rest of the world through trade. “To explore food security effects, IMPACT projects the percentage and number of malnourished preschool children (0 to 5 years old) in developing countries as a function of average per capita calorie availability, the share of females with secondary schooling, the 161 ratio of female to male life expectancy at birth, and the percentage of the population with access to safe water. A wide range of factors with potentially significant impacts on future developments in the world food situation can be modeled based on IMPACT. They include: population and income growth, the rate of growth in crop and livestock yield and production, feed ratios for livestock, agricultural research, irrigation and other investments, commodity price policies, and elasticities of supply and demand. For any specification of these underlying factors, IMPACT generates projections for crop area, yield, production, demand for food, feed and other uses, prices, and trade; and for livestock numbers, yield, production, demand, prices, and trade”. Developed at International Food Policy Research Institute. Macro-PDM Global scale river runoff model designed for use with IPCC scenarios. Changes in runoff are largely driven by change in precipitation, offset by increases in evaporation. Each 0.5 degree cell is treated as an independent catchment. Developed at Tyndall Centre. NEWS model suite - Global Nutrient Export from Watersheds Simulates river nutrient export from watersheds. “The Global NEWS system includes models that can be used to predict export of sediments, DIN, DIP, DOC, DON, dissolved organic P (DOP); particulate organic C (POC), PN, and PP” — based on factors such as inputs of fertiliser, manure and sewage. Developed through a workgroup of UNESCO’s International Oceanographic Commission. WaterGAP “The overall aim is to investigate current and future world-wide water availability, water use and water quality. Whereas water availability and water use have been implemented, a water quality module is currently under development and constitutes the next major goal. In particular, Water-GAP is concerned with the various impacts of global change on water availability and water demand, and to determine the development of water stress conditions on different spatial and temporal scales. Further research subjects are the variations of the water balance components on the hydrological large scale, and the future development of extreme conditions - such as floods and droughts”. Developed at University of Kassel / University of Frankfurt. 4.17.1 Participants Author: Lera Miles (UNEP — WCMP, Cambridge, United Kingdom) 162 4.18 Prioritisation of recommendations for future work This section builds from the results of the thematic reviews (sections 4.1 to 4.16) to prioritise amongst the analyses recommended those that are most important and most feasible for Phase 2. Scenario development The first priority for Phase 2, underlying all subsequent work, is the development of appropriate scenarios, built as counterfactuals to test the specific policy package that is being tested. As emphasised in sections 2.3.2 and 4.17, it is fundamental that in the evaluation of the economic consequences of biodiversity loss and ecosystem degradation, two scenarios are compared. These need to be identical in everything else but the specific policy package being tested. Otherwise, the economic results cannot be directly attributed to a difference in the state of biodiversity/ecosystems. Key steps in scenario development include spatially- explicit quantitative descriptions of the likely state of biodiversity under business-as-usual and under a set of policy interventions deemed to be sufficient to conserve biodiversity; and the identification and costing of those interventions. Care must be taken to ensure that the quantitative, mapped descriptions of the alternative states of the world are fit-for-purpose — that is to say that they address all of the key biophysical and socioeconomic input variables needed to feed into the models of the production and flow of each process and service of interest. Prioritising recommendations for Phase 2 The thematic reviews (sections 4.1 to 4.16) present an assessment of the feasibility, for each particular theme, of quantifying the provision of ecosystem processes or benefits in a spatially-explicit way as a basis for the economic valuation in Part 2. Recommendations are presented for how such quantification can be done, and the resources estimated to be necessary for each analysis. Here, we prioritise amongst those recommended analyses, based on two criteria (Figure 17): - Importance to human wellbeing: The degree to which these processes/benefits are likely to affect the overall results of the Review. This is based on a qualitative assessment of the predicted magnitude of the overall economic value of each process or benefit. Recommended analyses were coded, in decreasing order of importance, as A, B, or C. We have marked with an asterisk (*) those benefits that are particularly valuable for local livelihoods (which may have low overall economic value but may directly influence the wellbeing of millions of people). - Feasibility in Phase 2: The likelihood that the particular analysis recommended can be successfully undertaken in one year. Recommendations were coded, in decreasing order of feasibility, as 1, 2 or 3. The combination of these two criteria produces an overall priority ranking (Figure 17), from very high priority, when the analysis is both highly important and highly feasibly (A1), to very low priority, when the analysis is of lower importance and not particularly feasible in Phase 2 (C3). 163 importance to human wellbeing higher <————_—__ lower Very high priority ener 6s inPhase? | 2 ||‘ x SE tow rny 3 O x xX x Very low priority ike Figure 17 Prioritisation of future research lines according to their importance to human wellbeing (A to C) and feasibility in Phase 2 (1 to 3). Priority levels range from very high priority (A1) to very low priority (C3). higher High priority To do if resources available Table 5 presents the results of the prioritisation for each of the analyses recommended for the benefits/processes considered in this project. Prioritising recommendations for research in the longer-term For the longer-term, priorities are defined based on importance alone. Indeed, all of the analyses proposed are feasible, and so should all be done if time is less of a limitation. In this case, the focus should become to ensure that the most important benefits or processes are covered first. The priority then becomes to tackle analyses coded as A first, followed by B and by C. Table 5 Prioritisation of the recommended analyses for each category of benefit or process considered in this review. Benefit/process | Recommended analysis (see corresponding | Priority | Priority section for more detail) Scenario development Essential | Essential Wild crop Global pollination model building from pollination landscape-scale assessments Biological Global crop biological control model building control in crops from landscape-scale assessments 164 Genetic diversity | Global risk assessment of crop/livestock disease | B2?-3? of crops and from loss of genetic diversity livestock Soil quality for e Global valuation of improvement in soil quality | B3 crop production from changes in soil biota (internal effects) e Global valuation of soil subsidies when | B1* agriculture expands into natural ecosystems (conversion effects) Global model of the protective effects to crop | C2* soils of neighbouring natural habitat (neighbouring effects) Global model of value of wildlife fertilisers. C3* Livestock e Global model of rangeland contribution to | BI* livestock production Marine fisheries e Global model of marine fisheries provision Al* Inland fisheries e Global model of inland fisheries provision | B2* including aquaculture Wild animal Bl* td a Ma e Pantropical model of wild meat provision e European and North-American model of domestic recreational hunting products EF co © Fresh water e Global hydrological model for water provision | Al* provision and (quantity) regulation © Global hydrological model for water regulation | A2* (timing) Global hydrological model for water | A2* purification (quality) Nature-related e Global model of tourism in protected areas B3 outdoor activities e Global model of the value of green areas for | B2 local recreation 165 Global climate e Global terrestrial model of carbon storage and regulation sequestration 4.18.1 Participants Authors Ana Rodrigues, Andrew Balmford (University of Cambridge, UK) Matt Walpole (UNEP-WCMC, Cambridge, UK) 166 5 INVENTORY OF RESEARCH ORGANIZATIONS, PROGRAMS AND RECENT LITERATURE DEALING WITH ECONOMICS OF BIODIVERSITY LOSS As a contribution to the Scoping the Science project, an inventory was made of the main current research programs, organizations and literature dealing with economics of biodiversity loss. This Chapter presents the main results of these reviews and Annexes 6 and 7 list more detailed information on the results. It must be stressed that both the literature search and inventory of organisations and programs is not exhaustive, thus to obtain a more complete overview it is foreseen that this identification tasks will be continued as a part of the upcoming EU initiatives on economic of biodiversity loss. 5.1 Literature review An inventory was made of the main publications since 2005 (i.e. post the release of the Millennium Ecosystem Assessment). The research has been realised using the following search engines: Science direct, Scopus and Google scholar (succinctly), as well as direct search on the home pages of the Journals Ecological Economics, Nature and Science. A combination of key words was used during the research for each search engine. These included “Biodiversity + Economic Loss”, “Biodiversity + Economic Cost” and “Biodiversity + Economic Valuation”. From these key words the articles including all the terms in the text and published between 2005-2008 (included) have been selected. A total of 132 documents have been gathered, showing an increasing trend: 2005: 29 articles, 2006: 36 articles, 2007: 46 articles, 2008: 13 articles published and 8 articles in press. The selected articles come from different sources the main ones (more than 5 articles) being: Ecological economics (21), Science (14), Biological Conservation (10), and Conservation Biology (6). 5.2 Review of organisations, programmes and networks The inventory of organizations, programs and projects was carried out by using the main well-known websites as a starting point (see box 1). Most of these websites have a section with ‘links’ and by following these in a systematic way it is possible to obtain a good overview of the main organisations and programs. Similar to the literature search, keywords used were the link between “biodiversity” and “economics / economic loss / economic cost / economic valuation”. When going “deeper” into the various organisations and programs, many links are found to more specific subjects (e.g. programs, projects and organisations focusing on specific ecosystems (e.g. watershed- services, coral reefs, forests, etc), services or regions). Within the timeframe of this project it was not possible to explore these links in any detail but that could be possibly recommended as one of the future follow-up exercises. 167 BOX A. Main networks and websites dealing with economics of biodiversity (loss) Biodiversity Economics (www.biodiversityeconomics.org) Ecosystem Service Database (ESD) (//esd.uvm/edu) Ecosystem Services (www.esa.org/ecoservices) Ecosystem Services Database (esd-worldbank.org/eei/) Ecosystem Services project (www.ecosystemservicesproject.org) Ecosystem Valuation (www.ecosystemvaluation.org ) EnValue: www.epa.nsw.gov.au/envalue/ EVRI (www.evri.ca) MILLENNIUM ECOSYSTEM ASSESSMENT (www.maweb.org) Natural Capital Project (www.naturalcapitalproject.org) Nature Valuation & Financing Network (NV&F) (www.naturevaluation.org) In Annex 6 an overview is given of the main organisations and programs, organised by type of organisation including the following categories: main networks/websites; multi-lateral institutions, programs, treaties etc.; government supported initiatives; NGO’s; universities, research organizations & programs; and business supported initiatives. Per category a distinction is made between international and national organisations and programs. Within these categories, organizations are listed in alphabetical order for easy reference. This overview does not aim to be exhaustive, however in combination with the links given in Box A on the main networks & websites it should enable quick and easy access to the main organisations and programs on this topic’. In the future, some resources could be directed to identifying relevant networks at the national level. Ideally, a formal “network- analysis” should be performed to obtain more quantitative information on the main “nodes” in the national and international networks. One aim of this inventory has been to establish active links and working relations between the Scoping the Science project and the main organisations. Facilitating communication among the main Networks is an important objective of the Nature valuation and Financing "In case you miss an (important) organization or program, please send an email to do/f.degroot@wur.nl 168 Network (NV&F) (www.naturevaluation.org) which already has a quite extensive list of links and active working relations and could possibly be used to facilitate this process in the future. 169 6 REFERENCES Agardy, T., Alder, J., Dayton, P., Curran, S., Kitchingman, A., Wilson, M., Catenazzi, A., Restrepo, J., Birkeland, C., Blaber, S., Saifullah, S., Branch, G., Boersma, D., Nixon, S., Dugan, P., et al. 2005. Coastal Systems. In Millennium Ecosystem Assessment. Ecosystems and Human Well-being: Current States and Trends. Washington D.C., USA: World Resources Institute (pp. 515-543). Alcamo, J., Déll, P., Henrichs, T., Kaspar, F., Lehner, B., Résch, T., & Siebert, S. 2003. Development and testing of the WaterGAP 2 global model of water use and availability . Hydrological Sciences Journal 48: 317-337. Alcamo, J., Flérke, M., & Marker, M. 2007. Future long-term changes in global water resources driven by socio-economic and climatic changes / Changements futurs a long terme dans les ressources en eau globales forces par les changements climatiques et socio-economiques. Hydrological Sciences Journal 52: 247-275. Alcamo, J., Vuuren, D. V., Ringler, C., Cramer, W., Masui, T., Alder, J., & Schulze, K. 2005. Changes in nature’s balance sheet: model-based estimates of future worldwide ecosystem services. Ecology and _ Society 10: 19. [online] URL: http://www.ecologyandsociety.crg/voll 0/iss2/art19/. Alder, J., Guénette, S., Beblo, J., Cheung, W., & Christensen, V. in press. Ecosystem-based global fishing policy scenarios. Fisheries Centre Research Reports. Alfsen, K. H., Bye, T., Glomsrad, S., & Wiig, H. 2001. Soil degradation and economic development in Ghana. Environment and Development Economics 2: 119-143. Alin, S. R., Cohen, A. S., Bills, R., Gashagaza, M. M., Michel, E., Tiercelin, J., Martens, K., Coveliers, P., Mboko, S. K., West, K., Soreghan, M., Kimbadi, S., & Ntakimazi, G. 296. Effects of Landscape Disturbance on Animal Communities in Lake Tanganyika, East Africa. Conservation Biology 13: 1017-1033. Alkemade, R., Bakkenes, M., Bobbink, R., Miles, L., Nellemann, C., Simons, H., Tekelenburg, T., 2006. GLOBIO 3: Framework for the assessment of global terrestrial biodiversity. In: Bouwman, A. F., Kram, T., Klein (Eds.), 2006. Integrated modelling of global environmental change. An overview of IMAGE 2.4. Netherlands Environmental Assessment Agency, Bilthoven, the Netherlands, pp. 171-186. Allan, D., & Flecker, A. S. 1993. Biodiversity conservation in running waters. BioScience 43: 32-43. Allan, J. D., Abell, R., Hogan, Z., Revenga, C., Taylor, B.W., Welcomme, R. L. & Winemiller, K. 2005 Overfishing of Inland Waters. BioScience 55: 1041-1051 . Alongi, D. M. 2008. Mangrove forests: Resilience, protection from tsunamis, and responses to global climate change. Estuarine, Coastal and Shelf Science 76: 1-13. Alves, R. R., & Rosa, I. L. 2005. Why study the use of animal products in traditional medicines? Journal of Ethnobiology and Ethnomedicine 1. 170 Amornsakchai, S., Annez, P., Vongvisessomjai, S., Choowaew, S., Thailand Development Research Institute (TDRI), Kunurat, P., Nippanon, J., Schouten, R., Sripapatrprasite, P., Vaddhanaphuti, C., Vidthayanon, C., Wirojanagud, W., Watana, E. 2000. Pak Mun Dam, Mekong River Basin, Thailand. A WCD Case Study prepared as an input to the World Commission on Dams, Cape Town, South Africa: World Commission on Dams. Arnell, N. W. 1999. A simple water balance model for the simulation of streamflow over a large geographic domain. Journal of Hydrology 217: 314-335. Arnell, N. W. 2004. Climate change and global water resources: SRES emissions and socio-economic scenarios. Global Environmental Change 14: 31-52. Arnell, N. W., 2003. Effects of IPCC SRES emissions scenarios on river runoff: a global perspective. Hydrology and Earth System Sciences 7 (5), 619-641. Armold, J. G., & Fohrer, N. 2005. SWAT2000: current capabilities and research opportunities in applied watershed modelling. Hydrological Processes 19: 563-572. Arthurton, R., Barker, S., Rast, W., Huber, M., Alder, J., Chilton, J., Gaddis, E., Pietersen, K., Zockler, C., Al-Droubi, A., Dyhr-Nielsen, M., Finlayson, M., Fortnam, M., Kirk, E., Heileman, S., Rieu-Clark, A., et al. 2007. Chapter 4 - Water. In Global Environment Outlook 4. Nairobi, Kenya: GEO4 United Nations Environment Programme. Aylward, B. 2005. Land use, hydrological function and economic valuation. In M. Bonell & Leendert Adriaan Bruijnzeel (Eds.), Forests, Water, and People in the Humid Tropics: Past, Present, and Future. Cambridge, UK: Cambridge University Press (pp. 99-120). Badola, R., & Hussain, S. A. 2005. Valuing ecosystem functions: an empirical study on the storm protection function of Bhitarkanika mangrove ecosystem. Environmental Conservation 32: 85-92. Baillie, J., Hilton-Taylor, C., & Stuart, S. N. 2004. IUCN Red List of Threatened Species: A Global Species Assessment. Gland, Switzerland: IUCN. Baillie, J.E.M, Hilton-Taylor, C, Stuart, S.N (Eds.) 2004. 2004 IUCN Red List of Threatened Species. A Global Species Assessment. Gland, Switzerland and Cambridge UK: IUCN — The World Conservation Union. Baird, A. H., & Kerr, A. M. 2007. The Dangerous Myth of Tsunami Greenbelts: respose to Stone 2006. Science: E-Letter 314 Online: http://www.sciencemag.org/cgi/eletters/3 14/5798/404#5422. Baird, A., & Kerr, A. 2008. Landscape analysis and tsunami damage in Aceh: comment on Iverson and Prasad (2007). Landscape Ecology 23: 3-5. Bale, J., van Lenteren, J., & Bigler, F. 2008. Biological control and sustainable food production. Philosophical Transactions of the Royal Society B: Biological Sciences 363: 761-776. 171 Balirwa, J. S. 2007. Ecological, environmental and socioeconomic aspects of the Lake Victoria's introduced Nile perch fishery in relation to the native fisheries and the species culture potential: lessons to learn. African Journal of Ecology 45: 120-129. Barber, R. G., & Diaz, O. 1994. Maintenance of yields and soil fertility in nonmechanized cropping systems, bolivia. Soil Sci Soc Am J 58: 858-866. Barbier, E., Koch, E., Silliman, B., Hacker, S., Wolanski, E., Primavera, J., Granek, E., Polasky, S., Aswani, S., Cramer, L., Stoms, D., Kennedy, C., Bael, D., Kappel, C., Perillo, G., et al. 2008. Coastal Ecosystem-Based Management with Nonlinear Ecological Functions and Values. Science 319: 321-323. Bardgett, R. D., & McAlister, E. 1999. The measurement of soil fungal: bacterial biomass ratios as an indicator of ecosystem self-regulation in temperate meadow grasslands. Biology and Fertility of Soils 29: 282-290. Barnes, R. F. 2002. The bushmeat boom and bust in West and Central Africa. Oryx 36: 236-242. Bartholome, E., & Belward, A. 2005. GLC2000: a new approach to global land cover mapping from Earth observation data. International Journal of Remote Sensing 26: 1959-1977. Beck M. W, Heck Jr K. L, Able K. W, Childers D. L, Eggleston D. B, Gillanders B. M, Halpern B, Hays C. G, Hoshino K, Minello T. J, Orth R. J, Sheridan P. F, & Weinstein M. P. 2001. The identification, conservation, and management of estuarine and marine nurseries for fish and invertebrates. BioScience 51: 633-641. Bellwood, D. R., Hughes, T. P., Folke, C., & Nystrom, M. 2004. Confronting the coral reef crisis. Nature 429: 827-833. Bengtsson, J., Ahnstrom, J., & Weibull, A. 2005. The effects of organic agriculture on biodiversity and abundance: a meta-analysis. Journal of Applied Ecology 42: 261-269. Bennett, E. L. 2002. Is there a link between wild meat and food security? Conservation Biology 16: 590-592. Bennett, E.L. & Rao, M. 2002. Wild meat consumption in Asian tropical forest countries: is this a glimpse of the future for Africa? In S. Mainka & M. Trivedi (Eds) Links between Biodiversity, Conservation, Livelihoods and Food Security: The Sustainable Use of Wild Species for Meat. Gland, Switzerland: [UCN — The World Conservation Union (pp. 39-44). Bianchi, F., Booij, C., & Tscharntke, T. 2006. Sustainable pest regulation in agricultural landscapes: a review on landscape composition, biodiversity and natural pest control. Proceedings of the Royal Society B: Biological Sciences 273. Biesmeijer, J. C., Roberts, S. P. M., Reemer, M., Ohlemuller, R., Edwards, M., Peeters, T., Schaffers, A. P., Potts, S. G., Kleukers, R., Thomas, C. D., Settele, J., & Kunin, W. E. 2006. Parallel Declines in Pollinators and Insect-Pollinated Plants in Britain and the Netherlands. Science 313: 351-354. Bioversity International. 2008. Bioversity International - Plants and Animals. Online: http://www. bioversityinternational.org/Plants_and_Animals/index.asp (accessed 22 April 2008). Bird, W. 2005. Natural Fit: Can Green Space and Biodiversity Increase Levels of Physical Activity? Sandy, UK: RSPB. Bird, W. 2007. Natural Thinking. Sandy, UK: RSPB Bloem J., Hopkins, D.W. & Benedetti, A. (Eds.) 2006. Microbiological methods for assessing soil quality. Wallingford, UK: CABI. Bloem, J., de Ruiter, P.-C. & Bouwman, L.A. 1997. Food webs and nutrient cycling in agro- ecosystems. In J.D. van Elsas, J.T. Trevors and E. Wellington (Eds.), Modern Soil Microbiology. New York, USA: Marcel Dekker Inc (pp. 245-278). Bloem, J., Schouten, T., Didden, W., Jagers op Akkerhuis, G., Keidel, H., Rutgers, M., & Breure, T. 2004. Measuring soil biodiversity: experiences, impediments and research needs. In R. Francaviglia (Ed.), Agricultural impacts on soil erosion and soil biodiversity. developing indicators for policy analysis. Proceedings of the OECD expert meeting on soil erosion and soil biodiversity indicators, 25-28 March 2003, Rome, Italy. Paris, France: OECD. Bodmer, R. E., Eisenberg, J. F., & Redford, K. H. 1997. Hunting and the Likelihood of Extinction of Amazonian Mammals. Conservation Biology 11: 460-466. Bérjeson, L., Héjer, M., Dreborg, K.-H., Ekvall, T., Finnveden, G. 2005. Towards a user's guide to scenarios - a report on scenario types and scenario techniques. Royal Institute of Technology, Stockholm, Sweden. Botsford, L. W., Micheli, F., & Hastings, A. 2003. Principles for the design of marine reserves. 13: 25. Boumans, R., Costanza, R., Farley, J., Wilson, M. A., Portela, R., Rotmans, J., Villa, F., Grasso, M., 2002. Modeling the dynamics of the integrated earth system and the value of global ecosystem services using the GUMBO model. Ecological Economics 41 (3), 529-560. Bouwman, A.F., Kram, T., Klein (Eds.), 2006. Integrated modelling of global environmental change. An overview of IMAGE 2.4. Netherlands Environmental Assessment Agency, Bilthoven, the Netherlands. Bradshaw, C. J. A., Sohdi, N., Peh, K., & Brook, B. 2007. Global evidence that deforestation amplifies flood risk and severity in the developing world. Global Change Biology 13: 2379-2395. Brainard, J., Bateman, I., & Lovett, A. 2001. Modelling demand for recreation in English woodlands. Forestry 74: 423-438. Brander, K. M. 2007. Global fish production and climate change. Proceedings of the National Academy of Sciences USA 104:19709-19714. 173 Brashares, J. S., Arcese, P., Sam, M. K., Coppolillo, P. B., Sinclair, A. R. E., & Balmford, A. 2004. Bushmeat hunting, wildlife declines, and fish supply in West Africa. Science 306: 1180-1183. Brauman, K. A., Daily, G. C., Duarte, T. K., & Mooney, H. A. 2007. The nature and value of ecosystem services: an overview highlighting hydrologic services. Annual Review of Environment and Resources 32: 67-98. Brosi, B. J., Daily, G. C., Shih, T. M., Oviedo, F., & Duran, G. 2007 The effects of forest fragmentation on bee communities in tropical countryside. Journal of Applied Ecology. doi:10.1111/).1365-2664.2007.01412.x. Brown, C., & Grant, M. 2005. Biodiversity and Human Health: What Role for Nature in Healthy Urban Planning? Planning Healthy Towns and Cities 31: 326-338. Bruijnzeel, L. A. 2004. Hydrological functions of tropical forests: not seeing the soil for the trees? Agriculture, Ecosystems & Environment 104: 185-228. Bruijnzeel, L., van Dijk, A. I., van Noordwijk, M., Chappell, N. A., & Schellekens, J. 2007. Tropical deforestation, people and _ flooding. CIFOR. [online] http://www.worldagroforestry.org/water/downloads/bca_bruijnzeel.pdf. Brussaard, L., de Ruiter, P. C., & Brown, G. G. 2007. Soil biodiversity for agricultural sustainability. Agriculture, Ecosystems & Environment 121: 233-244. Bubb, P., May, I., Miles, L., & Sayer, J. 2004. The Cloud Forest Agenda Report. Cambridge, UK: UNEP-World Conservation Monitoring Centre. Bullock, A., & Acreman, M. 2003. The role of wetlands in the hydrological cycle. Hydrology and Earth System Sciences 7: 358-389. Carlton, J. T., Geller, J. B., Reaka-Kudla, M. L., Norse, E. A. 1999. Historical Extinctions in the Sea. Annual Review of Ecology and Systematics 30: 515-38. Carpenter, A. I., Marcus Rowcliffe, J., & Watkinson, A. R. 2004. The dynamics of the global trade in chameleons. Biological Conservation 120: 291-301. Carpenter, S. R., & Kitchell, J. F. 1988. Consumer control of lake productivity. BioScience 38: 764-769. Carpenter, S. R., Kitchell, J. F., Hodgson, J. R., Cochran, P. A., Elser, J. J., Elser, M. M., Lodge, D. M., Kretchmer, D., He, X., & von Ende, C. N. 1987. Regulation of lake primary productivity by food web structure. Ecology 68: 1863-1876. Carpenter, S. R., Pingali, P. L., Bennett, E. M., Zurek, M. B. (Eds.), 2005. Ecosystems and Human Well-Being: Volume 2 Scenarios: Findings of the Scenarios Working Group. Island Press, Washington DC. Carr, D., Pan, W., & Bilsborrow, R. 2006. Declining fertility on the frontier: the Ecuadorian Amazon. Population & Environment 28: 17-39. 174 Carver, S., Evans, A., & Fritz, S. 2002. Wilderness attribute mapping in the United Kingdom. International Journal of Wilderness 8: 25-29. Cassman, K. G., Wood, S., Choo, P. S., Cooper, D. H., Devendra, C., Dixon, J., Gaskell, J., Khan, S., Lal, R., Lipper, L., Pretty, J., Primavera, J., Ramankutty, N., Viglizzo, E., Wiebe, K., et al. 2005. Cultivated Systems. In Millennium Ecosystem Assessment. Ecosystems and Human Well-being: Current States and Trends. Washington D.C., USA: Island Press (pp. 745-794). Cassman, K. G., Wood, S., Choo, P. S., Cooper, H. D., Devendra, C., Dixon, J., Gaskell, J., Khan, S., Lal, R., Lipper, L., Pretty, J., Primavera, J.. Ramankutty, N., Viglizzo, E., Wiebe, K., et al. 2005. Chapter 26: Cultivated Systems. In Millennium Ecosystem Assessment. Current State & Trends Assessment. Washington DC, USA: Island Press (pp. 745-794). Ceballos-Lascurain, H. 1996. Tourism, ecotourism and protected areas. \UCN, Gland, Switzerland. Chamberlain, A. 1910. Scenery as a national asset. Outlook 95: 169. Chatenoux, B. & Peduzzi, P.. 2007. Impacts from the 2004 Indian Ocean Tsunami: analysing the potential protecting role of environmental features. Natural Hazards 40: 289-304. Chatenoux, B., & Peduzzi, P. 2005. Analysis on the Role of Bathymetry and other Environmental Parameters in the Impacts from the 2004 Indian Ocean Tsunami. Geneva, Switzerland: UNEP/GRID-Europe. Retrieved April 14, 2008, from http://www. grid.unep.ch/product/publication/download/environment_impacts_tsunam i.pdf Chivenge, P., Murwira, H., Giller, K., Mapfumo, P., & Six, J. 2007. Long-term impact of reduced tillage and residue management on soil carbon stabilization: Implications for conservation agriculture on contrasting soils. Soil and Tillage Research 94: 328-337. Christensen, V., & Walters, C. J. 2004. Ecopath with Ecosim: methods, capabilities and limitations. Ecological Modelling 172: 109-139. Christensen, V., & Walters, C. J. 2005. Using ecosystem modeling for fisheries management: Where are we? ICES Annual Science Conference, Aberdeen, Scotland, 20-24 September 2005: 1-16. Christensen, V., Guenette, S., Heymans, J. J., Walters, C. J., Watson, R., Zeller, D., & Pauly, D. 2003. Hundred-year decline of North Atlantic predatory fishes. Fish and Fisheries 4: 1-24. Christie, M., Han ley, N. & Hynes, S. 2007. Valuing enhancements to forest recreation using choice experiment and contingent behaviour models. Journal of Forest Economics 13: 75-102. Christner, B. C., Morris, C. E., Foreman, C. M., Cai, R., & Sands, D. C. 2008. Ubiquity of Biological Ice Nucleators in Snowfall. Science 319: 1214. LD CITES (Convention on International Trade in Endangered Species of Wild Fauna and Flora). 2008. The CITES Species. Convention on International Trade in Endangered Species of Wild Fauna and Flora. Retrieved April 26, 2008, from http://www.cites.org/eng/disc/species.shtml Claudet, J., Osenberg, C. W., Benedetti-Cecchi, L., Domenici, P., Garcia-Charton, J., Perez-Ruzafa, A., Badalamenti, F., Bayle-Sempere, J., Brito, A., Bulleri, F., Culioli, J., Dimech, M., Falcon, J. M., Guala, I., Milazzo, M., et al. in press. Marine reserves: size and age do matter. Ecology Letters. Clayton, L., Keeling, M., & Milner-Gulland, E. J. 1997. Bringing home the bacon: a spatial model of wild pig hunting in Sulawesi, Indonesia. Ecological Applications 7: 642- 652. Coates D., Ouch Poeu, Ubolratana Suntornratana, N Thanh Tung & Sinthavong Viravong. 2003. Biodiversity and fisheries in the Lower Mekong Basin. Mekong Development Series No. 2. Phnom Penh, Vietnam: Mekong River Commission. Cochard, R., Ranamukhaarachchi, S. L., Shivakoti, G. P., Shipin, O. V., Edwards, P. J., & Seeland, K. T. 2008. The 2004 tsunami in Aceh and Southern Thailand: A review on coastal ecosystems, wave hazards and vulnerability. Perspectives in Plant Ecology, Evolution and Systematics 10: 3-40. Cole, F. R., Medeiros, A. C., Loope, L. L., & Zuehlke, W. W. 1992. Effects of the Argentine ant on arthropod fauna of Hawaiian high-elevation shrubland. Ecology 73: 1313-1322. Collen, P. and Gibson, R. J. 2001 The general ecology of beavers (Castor spp.), as related to their influence on stream ecosystems and riparian habitats, and the subsequent effects on fish — a review. Reviews in Fish Biology and Fisheries 10: 439-461. Cook, R. M., Sinclair, A., & Stefansson, G. 1997. Potentiai collapse of North Sea cod stocks. Nature 385: 521-522. Cooke, S. J. & Cowx, I. G. 2006 Contrasting recreational and commercial fishing: Searching for common issues to promote unified conservation of fisheries resources and aquatic environments. Biological Conservation 128: 93-108. Corlett, R. T. 2007. The impact of hunting on the mammalian fauna of tropical Asian forests. Biotropica 39: 292-303. Costanza R., d'Arge R., Groot R.d., Farber S., Grasso M., Hannon B., Limburg K., Naeem S., O'Neill R.V., Paruelo J., Raskin R.G., Sutton P. & Belt M.v.d. (1997) The value of the world's ecosystem services and natural capital. Nature 387: 253 - 260. Costanza, R., 2000. Visions of alternative (unpredictable) futures and their use in policy analysis. Ecology and Society 4 (1), 5+ Costanza, R., Mitsch, W. J., & Day, J. 2006. Creating a sustainable and desirable New Orleans. Ecological Engineering 26: 317-320. 176 Costanza, R., Pérez-Maqueo, O., Martinez, M. L., Sutton, P., Anderson, S. J., & Mulder, K. in press. The value of coastal wetlands for hurricane protection. Ambio. Cote, I. M., Mosqueira, I., & Reynolds, J. D. 335. Effects of marine reserve characteristics on the protection of fish populations: a meta-analysis. Journal of Fish Biology 59: 178-189. Cowlishaw, G., Mendelson, S., & Rowcliffe, J. M. 2005. Evidence for post-depletion sustainability in a mature bushmeat market. Journal of Applied Ecology 42: 460-468. Crookes, D., Ankudey, N., & Milner-Gulland, E. J. 2005. The value of a long-term bushmeat market dataset as an indicator of system dynamics. Environmental Conservation 32: 333-339. Crooks, K. R., & Soulé, M. E. 1999. Mesopredator release and avifaunal extinctions in a fragmented system. Nature 400: 563-566. Daby, D. 2003. Effects of seagrass bed removal for tourism purposes in a Mauritian bay. Environmental Pollution 125: 313-324. Dahdouh-Guebas, F., Jayatissa, L., Di Nitto, D., Bosire, J., Lo Seen, D., & Koedam, N. 2005. How effective were mangroves as a defence against the recent tsunami? Current Biology 15: R443-R447. Daily, G., Ceballos, G., Pacheco, J., Suzan, G., & Sanchez-Azofeita, A. 2003. Countryside biogeography of Neotropical mammals: conservation opportunities in agricultural landscapes of Costa Rica. Conservation Biology 17: 1814-1826. Danielsen, F., Sorensen, M. K., Olwig, M. F., Selvam, V., Parish, F., Burgess, N. D., Hiraishi, T., Karunagaran, V. M., Rasmussen, M. S., Hansen, L. B., Quarto, A., & Suryadiputra, N. 2005. The Asian Tsunami: A Protective Role for Coastal Vegetation. Science 310: 643. Darby, S. E. 1999. Effect of riparian vegetation on flow resistance and flood potential. Journal of Hydraulic Engineering 125: 443-454. Darwall, W., Smith, K., Lowe, T. and Vié, J.-C. 2005. The Status and Distribution of Freshwater Biodiversity in Eastern Africa. Gland, Switzerland and Cambridge, UK: IUCN SSC Freshwater Biodiversity Assessment Programme. IUCN. Daskalov, G. M., Grishin, A. N., Rodionov, S., & Mihneva, V. 2007. Trophic cascades triggered by overfishing reveal possible mechanisms of ecosystem regime shifts. Proceedings of the National Academy of Sciences USA 104: 10518-10523. DeFoliart, G. R. 1999. Insects as food: why the western attitude is important. Annual review of entomology 44: 21-50. Devine, J. A., Baker, K. D., & Haedrich, R. L. 2006. Fisheries: Deep-sea fishes qualify as endangered. Nature 439: 29. Di Falco, S., & Perrings, C. 2005. Crop biodiversity, risk management and the implications of agricultural assistance. Ecological Economics 55: 459-466. 177 Diaz, S., Tilman, D., Fargione, J., Chapin, F.S., Dirzo, R., Kitzberger, T., Gemmill, B., Zobel, M., Vila, M., Mitchell, C., Wilby, A., Daily, G.C., Galetti, M., Laurance, W.F., Pretty, J., et al. 2005. Chapter 11: Biodiversity Regulation of Ecosystem Services. In Millennium Ecosystem Assessment: Current State & Trends Assessment. Washington DC, USA: Island Press (pp. 297-329). Dickie, I., Hughes, J., & Esteban, A. 2006. Watched like never before; the local economic benefits of spectacular bird species. Sandy, UK: RSPB. Dolfman, M. L., Wasser, S. F., & Bergman, B. 2007. The effects of Hurricane Katrina on the New Orleans economy. Monthly Labor Review: 3-18. Doll, P., & Siebert, S. 2002. Global modeling of irrigation water requirements. Water Resources Research 38: 1037. Dll, P., Kaspar, F., & Lehner, B. 2003. A global hydrological model for deriving water availability indicators: model tuning and validation. Journal of Hydrology 270: 105- 134. Donlan, J., Berger, J., Bock, C. E., Bock, J., Burney, D. A., Estes, J. A., Foreman, D.., Martin, P., Roemer, G., Smith, F., Soule, M., & Greene, H. 2006. Pleistocene Rewilding: An optimistic agenda for Twenty First Century conservation. The American Naturalist 168: 660-681. Douglas, E., Wood, S., Sebastian, K., Vorosmarty, C., Chomitz, K., & Tomich, T. 2007. Policy implications of a pan-tropic assessment of the simultaneous hydrological and biodiversity impacts of deforestation. Water Resources Management 21: 211-232. Duarte, C. M. 2002. Marine biodiversity and ecosystem services: an elusive link. Journal of Experimental Marine Biology and Ecology 250: 117-131. Dudley, N., & Stolton, S. 2003. Running pure: the importance of forest protected areas to drinking water. World Bank/WWF Alliance for Forest Conservation and Sustainable Use. Dugan, P., Dey, M. M. and Sugunan, V. V. in press. Fisheries and water productivity in tropical river basins: Enhancing food security and livelihoods by managing water for fish. Dugan, P., Sugunan, V. V., Welcomme. R. L., Béné, C., Brummett, R. E., & Beveridge, M. C. 2007. Inland fisheries and aquaculture. In D. Molden (Ed.), Water for Food, Water for Life: A Comprehensive Assessment of Water Management in Agriculture. Washington D.C., USA: Earthscan and International Water Management Institute (pp. 459-483). Duraiappah, A. K., Naeem, S., Agardy, T., Ash, N. J., Cooper, H. D., Diaz, S., Faith, D. P., Mace, G., McNeely, J. A., Mooney, H. A., Oteng-Yeboah, A. A., Pereira, H. M., Polasky, S., Prip, C., Reid, W. V., et al. 2005. Biodiversity Synthesis Report. In Millennium Ecosystem Assessment: Synthesis Reports. Washington DC, USA: Island Press. 178 Eagles, P. 2002. Trends in park tourism: economics, finance and management. Journal of Sustainable Tourism 10: 132-153. Eagles, P.F.J., McLean, D. & Stabler, M.J. 2000. Estimating the tourism volume and value in parks and protected areas in Canada and the USA. George Wright Forum 17, 62- 82. East, T., Kiimpel, N. F., Milner-Gulland, E., & Rowcliffe, J. M. 2005. Determinants of urban bushmeat consumption in Rio Muni, Equatorial Guinea. Biological Conservation 126: 206-215. EJF (Environmental Justice Foundation). 2006. Mangroves: Nature’s defence against Tsunamis— A report on the impact of mangrove loss and shrimp farm development on coastal defences. London, UK: Environmental Justice Foundation. Retrieved March 27, 2008, from http://www.ejfoundation.org/pdf/tsunami_report.pdf Elliott, J., Grahn, R., Sriskanthan, G., & Arnold, C. 2002. Wildlife and Poverty Study. Livestock and Wildlife Advisory Group, Department for International Development. Retrieved April 28, 2008, from http://www.iwmce.org/IWMC- Forum/Articles/WildlifePovertyStudy.pdf Enright, W. D. 2000. The effect of terrestrial invasive alien plants on water scarcity in South Africa. Physics and Chemistry of the Earth, Part B: Hydrology, Oceans and Atmosphere 25: 237-242. Entwistle, A. & Dunstone, N. 2000. Priorities for the conservation of mammalian biodiversity. Cambridge University Press, Cambridge, UK. Ernst, C. 2004. Protecting the Source: Land Conservation and the Future of America’s Drinking Water. Washington, D.C.: Trust for Public Land. EuroCoML 2008. Census of Marine Life on Seamounts. Available online: http://www.eurocoml.org/index.php?option=com_content& task=view&id=83 & Itemid =119. Accessed March 3 2008. Fa, J. E., & Purvis, A. 1997. Body size, diet and population density in afrotropical forest mammals: a comparison with neotropical species. The Journal of Animal Ecology 66: 98-112. Fa, J. E., Currie, D., & Meeuwig, J. 2003. Bushmeat and food security in the Congo Basin: linkages between wildlife and people's future. Environmental Conservation 30: 71-78. Fa, J. E., Peres, C. A., & Meeuwig, J. 2002. Bushmeat exploitation in tropical forests: an intercontinental comparison. Conservation Biology 16: 232-237. Fa, J. E., Ryan, S. F., & Bell, D. J. 2005. Hunting vulnerability, ecological characteristics and harvest rates of bushmeat species in afrotropical forests. Biological Conservation 121: 167-176. Fa, J. E., Seymour, S., Dupain, J., Amin, R., Albrechtsen, L., & Macdonald, D. 2006. Getting to grips with the magnitude of exploitation: Bushmeat in the Cross-Sanaga rivers region, Nigeria and Cameroon. Biological Conservation 129: 497-510. 179 Fa, J.E. & Peres, C. 2001. Game vertebrate extraction in African and Neotropical forests: an intercontinental comparison. In J. Reynolds, G. Mace, J.G. Robinson and K. Redford (Eds). Conservation of Exploited Species. Cambridge, UK: Cambridge University Press. (pp. 203-241). FACE (Federation of Associations for Hunting and Conservation of the EU). 2007. Census of the number of hunters in Europe. Brussels, Belgium: FACE: Federation of Associations for Hunting and Conservation of the EU. Retrieved April 25, 2008, from http://www.face-europe.org/fs-hunting.htm Fahrig, L. 2003. Effects of habitat fragmentation on biodiversity. Annual Review of Ecology, Evolution, and Systematics 34: 487-515. FAO (Food and Agriculture Organization) & CIFOR (Center for International Forestry Research). 2005. Forests and floods. Drowning in fiction or thriving on facts? Bogor Barat, Indonesia: Center for International Forestry Research and Food and Agriculture Organization of the United Nations. Retrieved April 14, 2008, from http://www. fao.org/docrep/008/ae929e/ae929e00.htm FAO (Food and Agriculture Organization). 1999. Sustaining agricultural biodiversity and agro-ecosystem functions: Opportunities, incentives and approaches for the conservation and sustainable use of agricultural biodiversity in agro-ecosystems and production systems. Rome, Italy: Food and Agriculture Organization of the United Nations. FAO (Food and Agriculture Organization). 2006. The State of the World's Fisheries and Aquaculture 2006. Rome, Italy: Food and Agriculture Organization of the United Nations (FAO). FAO. 1997. The State of the World’s Plant Genetic Resources for Food and Agriculture. Rome, Italy: Food and Agriculture Organization of the United Nations. FAO. 2007. The State of the World’s Animal Genetic Resources for Food and Agriculture (B. Rischkowsky & D. Pilling, Eds.). Rome, Italy: Food and Agriculture Organization of the United Nations. Ferreras, P., Gaona, P., Palomares, F., Delibes, M., 2001. Restore habitat or reduce mortality? implications from a population viability analysis of the Iberian lynx. Animal Conservation 4 (3), 265-274. Ferreras, P., Gaona, P., Palomares, F., Delibes, M., 2001. Restore habitat or reduce mortality? implications from a population viability analysis of the Iberian lynx. Animal Conservation 4 (3), 265-274. Finlayson, C. M., D’Cruz, R., Aladin, N., Barker, D. R., Beltram, G., Brouwer, J., Davidson, N., Duker, L., Junk, W., Kaplowitz, M. D., Ketelaars, H., Kreuzberg- Mukhina, E., Espino, G. D. L. L., Leveque, C., Lopez, A., et al. 2005. Inland Water Systems. In Millennium Ecosystem Assessment. Ecosystems and Human Well-being: Current States and Trends. Washington D.C., USA: Island Press (pp. 551-583). 180 Fischer G, Shah M, van Velthuizen H,Nachtergaele FO. 2002. Global Agroecological Assessment for Agriculture in the 21st Century: Methodology and Results. Laxenburg, Austria: International Institute for Applied Systems Analysis. Fisher B., Turner R.K., Balmford A., Burgess N.D., Green R., Kajembe G., Kulindwa K., Lewis S., Marchant R., Morse-Jones S., Naidoo R., Paavola J., Ricketts T. & Rouget M. (in review) Valuing the Arc: an ecosystem services approach for integrating natural systems and human welfare in the Eastern Arc Mountains of Tanzania. Fjortoft, I. 2001. The Natural Environment as a Playground for Children: The Impact of Outdoor Play Activities in Pre-Primary School Children. Early Childhood Education Journal 29: 111-117. Foley, J. A., DeFries, R., Asner, G. P., Barford, C., Bonan, G., Carpenter, S. R., Chapin, F. S., Coe, M. T., Daily, G. C., Gibbs, H. K., Helkowski, J. H., Holloway, T., Howard, E. A., Kucharik, C. J., Monfreda, C., et al. 2005. Global consequences of land use. Science 309: 570-574. Folke, C., Carpenter, S., Walker, B., Scheffer, M., Elmqvist, T., Gunderson, L., & Holling, C. S. 2004. Regime shifts, Resilience and Biodiversity in Ecosystem Management. Annual Review of Ecology, Evolution and Systematics 35: 557-581. Fossa, J. H., Mortensen, P. B., & Furevik, D. M. 2002. The deep-water coral Lophelia pertusa in Norwegian waters: distribution and fishery impacts. Hydrobiologia 471: 1- 2 Fragoso, J. M. V. 1997. Tapir-generated seed shadows: scale-dependent patchiness in the Amazon rain forest. The Journal of Ecology 85: 519-529. Fragoso, J. M. V., Silvius, K. M., & Correa, J. A. 2003. Long-distance seed dispersal by tapirs increases seed survival and aggregates tropical trees. Ecology 84: 1998-2006 . Frank, K. T., Petrie, B., Choi, J. S., & Leggett, W. C. 2005. Trophic cascades in a formerly cod-dominated ecosystem. Science 308: 1621-1623. Freiwald, A., Fossa, J. H., Grehan, A., Koslow, T., Roberts, J.M. (2004). Cold-water coral reefs: out of sight - no longer out of mind. Cambridge, UK: UNEP-WCMC. Fromentin, J., & Powers, J. E. 2005. Atlantic bluefin tuna: population dynamics, ecology, fisheries and management. Fish and Fisheries 6: 281-306. Gardner, T. A., Cote, I. M., Gill, J. A., Grant, A., & Watkinson, A. R. 2003. Long-term region-wide declines in Caribbean corals. Science 301: 958-960. Gerten, D., Schaphoff, S., Haberlandt, U., Lucht, W., & Sitch, S. 2004. Terrestrial vegetation and water balance--hydrological evaluation of a dynamic global vegetation model. Journal of Hydrology 286: 249-270. Ghazoul, J. 2005. Buzziness as usual? Questioning the global pollination crisis. Trends in Ecology & Evolution 20: 367-373. 181 Gilbert, F. F. 2000. Considerations in managing wildlife populations for sport. Wildlife Society Bulletin 28: 459-463. Giles-Corti, B., Broomhall, M. H., Knuiman, M., Collins, C., Douglas, K., Ng, K., Lange, A., & Donovan, R. J. 2005. Increasing walking: How important is distance to, attractiveness, and size of public open space? American Journal of Preventive Medicine 28: 169-176. Giller, K. & Palm, C. 2004. Cropping systems: slash-and-burn cropping systems of the tropics. In Encyclopedia of Plant and Crop Science. UK: CPL Scientific Publishing Services (pp. 363-366). Giller, K. E., Beare, M. H., Lavelle, P., Izac, A. -. N., & Swift, M. J. 1997. Agricultural intensification, soil biodiversity and agroecosystem function. Applied Soil Ecology 6: 3-16. Goodwin, H, Kent, I, Parker, K & Walpole, M. 1998. Tourism, conservation and sustainable development: case studies from Africa and Asia. ITED, London, UK. Goodwin, H. & Leader-Williams, N. 2000. Tourism and Protected Areas - distorting conservation priorities towards, charismatic megafauna? Chapter 14 in Entwistle, A & Dunstone N Priorities for the conservation of mammalian biodiversity. Cambridge University Press, Cambridge, UK. Goodwin, H.J. 1996. In pursuit of ecotourism. Biodiversity & Conservation 5: 277-291. Gordon, H., Haygarth, P. M., & Bardgett, R. D. 2008. Drying and rewetting effects on soil microbial community composition and nutrient leaching. Soil Biology and Biochemistry 40: 302-311. Gordon, L. J., Peterson, G. D., & Bennett, E. M. 2008. Agricultural modifications of hydrological flows create ecological surprises. Trends in Ecology & Evolution 23: 211-219. Gordon, L. J., Steffen, W., Jonsson, B. F., Folke, C., Falkenmark, M., & Johannessen, A. 2005. Human modification of global water vapor flows from the land surface. Proceedings of the National Academy of Sciences 102: 7612-7617. Gossling, S. 1999. Ecotourism: a means to safeguard biodiversity and ecosystem functions? Ecological Economics 29: 303-320. Goudarzi, S. 2006. Ape Meat Sold in U.S., European Black Markets. National Geographic News: July 18, 2006. [Online] http://news.national geographic.com/news/2006/07/060718-ape-meat.html. Goulson, D., Hanley, M., Darvill, B., Ellis, J., & Knight, M. 2005. Causes of rarity in bumblebees. Biological Conservation 122: 1-8. Goulson, D., Lye, G. C., & Darvill, B. 2008. Decline and conservation of bumble bees. Annual review of entomology 53: 191-208. 182 Gram, S. 2001. Economic valuation of special forest products: an assessment of methodological shortcomings. Ecological Economics 36: 109-117. Green, E. P., & Short, F. T. (Eds.). 2003. World Atlas of Seagrasses. Berkley, CA, USA: University of California Press. Retrieved April 17, 2008, from http://www.unep- wemce.org/marine/seagrassatlas/index.htm Green, R. E., Cornell, S. J., Scharlemann, J. P., Balmford, A., 2005. Farming and the fate of wild nature. Science 307 (5709), 550-555. Greenleaf, S. S., & Kremen, C. 2006. Wild bees enhance honey bees' pollination of hybrid sunflower. Proceedings of the National Academy of Sciences of the United States of America 103: 13890-5. Griffiths, B. S., Ritz, K., Bardgett, R. D., Cook, R., Christensen, S., Ekelund, F., Sorensen, S. J., Baath, E., Bloem, J., Ruiter, P. C. D., Dolfing, J., & Nicolardot, B. 2000. Ecosystem response of pasture soil communities to fumigation-induced microbial diversity reductions: an examination of the biodiversity - ecosystem function relationship. Oikos 90: 279-294. Groot, R. D., Ramakrishnan, P., Berg, A. V. D., Kulenthran, T., Muller, S., Pitt, D., Wascher, D., Wijesuriya, G., Amelung, B., Eliezer, N., Gopal, A. R., & Rossler, M. 2005. Cultural and Amenity Services. In Millennium Ecosystem Assessment. Ecosystems and Human Well-being: Current States and Trends. Washington D.C., USA: World Resources Institute (pp. 457-474). Gruber, U., & Bartelt, P. 2007. Snow avalanche hazard modelling of large areas using shallow water numerical methods and GIS. Environmental Modelling & Software 22: 1472-1481. Guenni, L. B., Cardoso, M., Goldammer, J., Hurtt, G., Mata, L. J., Ebi, K., House, J., Valdes, J., & Norgaa, R. 2005. Regulation of Natural Hazards: Floods and Fires. In Millennium Ecosystem Assessment. Ecosystems and Human Well-being: Current States and Trends. Washington D.C., USA: Island Press (pp. 441-454). Guidetti, P. & Sala, E. 2007. Community-wide effects of marine reserves in the Mediterranean Sea. Marine Ecology Progress Series 335: 43-56. Gullstré6m M, de la Torre C. M., Bandeira S.O., Bjérk M., Dahlberg M., Kautsky, N., R6nnbiack, P. and Ohman M. C. 2002. Seagrass Ecosystems in the Western Indian Ocean. AMBIO 31: 588-596. Halls A. & Welcomme R.L. (2004) Dynamics of river fish populations in response to hydrological conditions: A simulation study. River Research and Applications 20:985-1000. Halpern, B. S. 2003. The impact of marine reserves: do reserves work and does reserve size matter? Ecological Applications 13: 117. Halpern, B. S., Walbridge, S., Selkoe, K. A., Kappel, C. V., Micheli, F., D'Agrosa, C.., Bruno, J. F., Casey, K. S., Ebert, C., Fox, H. E., Fujita, R., Heinemann, D., Lenihan, 183 H. S., Madin, E. M. P., Perry, M. T., et al. 2008. A global map of human impact on marine ecosystems. Science 319: 948-952. Heard, M. S., Carvell, C., Carreck, N. L., Rothery, P., Osborne, J. L., & Bourke, A. F. G. 2007. Landscape context not patch size determines bumble-bee density on flower mixtures sown for agri-environment schemes. Biology Letters 3: 638-41. Heberlein, T., Ericsson, G., & Wollscheid, K. 2002. Correlates of hunting participation in Europe and North America. Zeitschrift fiir Jagdwissenschaft 48: 320-326. Henwood, K. 2003. Is there a role for environmental and countryside agencies in promoting benefits to health? London, UK: NHS National Institute for Health and Clinical Excellence. Hiddink, J., MacKenzie, B., Rijnsdorp, A., Dulvy, N., Nielsen, E., Bekkevold, D., Heino, M., Lorance, P., & Ojaveer, H. in press. Importance of fish biodiversity for the management of fisheries and ecosystems. Fisheries Research. Hilborn, R., Quinn, T. P., Schindler, D. E., & Rogers, D. E. 2003a. Biocomplexity and fisheries sustainability. Proceedings of the National Academy of Sciences 100: 6564- 6568. Hilborn, R., Trevor A. Branch, Billy Ernst, Arni Magnusson, Carolina V. Minte-Vera, Mark D. Scheuerell, A., & Juan L. Valero. 2003b. State of the world's fisheries. Annual Review of Environment and Resources 28: 359-399. Hill, J. O., & Peters, J. C. 1998. Environmental Contributions to the Obesity Epidemic. Science 280: 1371-1374. Hitzfeld, B. C., Héger, S. J., & Dietrich, D. R. 2000. Cyanobacterial toxins: removal during drinking water treatment, and human risk assessment. Environmental Health Perspectives 108. Hoegh-Guldberg, O., Mumby, P. J., Hooten, A. J., Steneck, R. S., Greenfield, P., Gomez, E., Harvell, C. D., Sale, P. F., Edwards, A. J., Caldeira, K., Knowlton, N., Eakin, C. M., Iglesias-Prieto, R., Muthiga, N., Bradbury, R. H., et al. 2007. Coral reefs under rapid climate change and ocean acidification. Science 318: 1737-1742. Hofer, D. 2002. The Lion's Share of the Hunt. Trophy Hunting and Conservation: a Review of the Legal Eurasian Tourist Hunting Market and Trophy Trade Under CITES. Brussels, Belgium: TRAFFIC Europe. Holzschuh, A., Steffan-Dewenter, I., & Tscharntke, T. 2008. Agricultural landscapes with organic crops support higher pollinator diversity. Oikos 117: 354-361. Holzschuh, A., Steffan-Dewenter, I., Kleijn, D., & Tscharntke, T. 2007. Diversity of flower-visiting bees in cerea! fields: effects of farming system, landscape composition and regional context. Journal of Applied Ecology 44: 41-49. Hooper, D. U., Chapin, F. S., Ewel, J. J., Hector, A., Inchausti, P.., Lavorel, S., Lawton, J. H., Lodge, D. M., Loreau, M., Naeem, S., Schmid, B., Setéla, H., Symstad, A. J., 184 Vandermeer, J., & Wardle, D. A. 2005. Effects of biodiversity on ecosystem functioning: A consensus of current knowledge. Ecological Monographs 75: 3-35 . House, J. I., Prentice, C. I., Le Quéré, C., 2002. Maximum impacts of future reforestation or deforestation on atmospheric CO2. Global Change Biology 8 (11), 1047-1052 House, J., Brovkin, V., Betts, R., Constanza, B., Dias, M. A. S., Holland, B., Quere, C. L., Phat, N. K., Riebesell, U., Scholes, M., Arneth, A., Barratt, D., Cassman, K., Christensen, T., Cornell, S., et al. 2005. Climate and Air Quality. In Millennium Ecosystem Assessment. Ecosystems and Human Well-being: Current States and Trends. Washington D.C., USA: Island Press (pp. 355-390). Hughes, B. B., Irfan, M. T., 2007. Assessing strategies for reducing poverty. The International Studies Review 9 (4), 690-710. Hughes, B.B., 2001. Choices in the face of uncertainty: the international futures (IFs) model. Futures 33 (1), 55-62. Hughes, T. P., Bellwood, D. R., Folke, C., Steneck, R. S., & Wilson, J. 2005. New paradigms for supporting the resilience of marine ecosystems. Trends in Ecology & Evolution 20: 380-386. Hutchings, J. A. & Reynolds, J. D. 2004 Marine fish population collapses: consequences for recovery and extinction risk. BioScience. 54: 297-309. Hutchings, J. A. 2000. Collapse and recovery of marine fishes. Nature 406: 882-885. IPCC. 2001. IPCC Special Report on Emissions Scenarios. Geneva, Switzerland: Intergovernmental Panel on Climate Change. IPCCWGI. 2007. Climate Change 2007 - The Physical Science Basis. Working Group I Contribution to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK: Cambridge University Press. IPCCWGII. 2007. Climate Change 2007 — Impacts, Adaptation and Vulnerability. Working Group II contribution to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK: Cambridge University Press. IPCCWGIII. 2007. Climate Change 2007 - Mitigation of Climate Change. Working Group III Contribution to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK: Cambridge University Press. Iriarte, J. A., & Jaksic, F. M. 1986. The fur trade in Chile: An overview of seventy-five years of export data (1910-1984). Biological Conservation 38: 243-253. ISDR. 2006. The 2004 Indian Ocean Tsunami: One year after. Disaster Reduction in Asia Pacific - ISDR Informs: 47-68. Islam, M., & Haque, M. 2004. The mangrove-based coastal and nearshore fisheries of Bangladesh: ecology, exploitation and management. Reviews in Fish Biology and Fisheries 14: 153-180. 185 IUCN (The World Conservation Union). 2007. IUCN Red List of Threatened Species. IUCN - The World Conservation Union. Retrieved April 26, 2008, from http://www. iucnredlist.org/ Iverson, L., & Prasad, A. 2007. Using landscape analysis to assess and model tsunami damage in Aceh province, Sumatra. Landscape Ecology 22: 323-331. Jackson, E., Rowden, A., Attrill, M., Bossey, S., & Jones, M. 200la. The importance of seagrass beds as a habitat for fishery species. Oceanography and Marine Biology 39: 269-303. Jackson, J. B. C., Kirby, M. X., Berger, W. H., Bjorndal, K. A., Botsford, L. W., Bourque, B. J., Bradbury, R. H., Cooke, R., Erlandson, J., Estes, J. A., Hughes, T. P., Kidwell, S., Lange, C. B., Lenihan, H. S., Pandolfi, J. M., et al. 2001b. Historical Overfishing and the Recent Collapse of Coastal Ecosystems. Science 293: 629-637. Jalale, T. 1993. National parks tourism in Malawi. MPhil thesis, University of Kent, Canterbury, Kent. Jerozolimski, A., & Peres, C. A. 2003. Bringing home the biggest bacon: a cross-site analysis of the structure of hunter-kill profiles in Neotropical forests. Biological Conservation 111: 415-425. Johansson, D., Azar, C., June 2007. A scenario based analysis of land competition between food and bioenergy production in the US. Climatic Change 82: 267-291. Johnson, R., 2007. Recent honey bee colony declines. Congressional Research Service Report for Congress. RL 33938. Washington DC, USA. Kannan, N., White, S., & Whelan, M. 2007. Predicting diffuse-source transfers of surfactants to surface waters using SWAT. Chemosphere 66: 1336-1345. Kannan, N., White, S., Worrall, F., & Whelan, M. 2007. Hydrological modelling of a small catchment using SWAT-2000 - Ensuring correct flow partitioning for contaminant modelling. Journal of Hydrology 334: 64-72. Karesh, W. B., Cook, R. A., Bennett, E. L., & Newcomb, J. 2005. Wildlife trade and global disease emergence. Emerging Infectious Diseases 11: 1000-1002. Kathiresan, K., & Rajendran, N. 2005. Coastal mangrove forests mitigated tsunami. Estuarine, Coastal and Shelf Science 65: 601-606. Kean, J., Wratten, S., Tylianakis, J., & Barlow, N. 2003. The population consequences of natural enemy enhancement, and implications for conservation biological control. Ecology Letters 6: 604-612. Kearns, C. A., Inouye, D. W., & Waser, N. M. 1998. Endangered Mutualisms: The Conservation of Plant-Pollinator Interactions. Annual Review of Ecology and Systematics 29: 83-112. Kerr, A. M., & Baird, A. H. 2007. Natural Barriers to Natural Disasters. BioScience 57: 102-103. 186 Kerr, A. M., Baird, A. H., & Campbell, S. J. 2006. Comments on "Coastal mangrove forests mitigated tsunami" by K. Kathiresan and N. Rajendran [Estuar. Coast. Shelf Sci. 65 (2005) 601-606]. Estuarine, Coastal and Shelf Science 67: 539-541. Kindermann, G. E., Obersteiner, M., Rametsteiner, E., McCallum, I., 2006. Predicting the deforestation-trend under different carbon-prices. Carbon Balance and Management nA): Klein, A., Cunningham, S. A., Bos, M. & Steffan-Dewenter, I. in press. Advances in Pollination Ecology from Tropical Plantation Crops. Ecology. Klein, A., Steffan-Dewenter, I., & Tscharntke, T. 2003a. Pollination of Coffea canephora in relation to local and regional agroforestry management. Journal of Applied Ecology 40: 837-845. Klein, A., Steffan-Dewenter, I., & Tscharntke, T. 2003b. Fruit set of highland coffee increases with the diversity of pollinating bees. Proceedings of the Royal Society B: Biological Sciences 270: 955-961. Klein, A., Steffan-Dewenter, I., & Tscharntke, T. 2003c. Bee pollination and fruit set of Coffea arabica and C. canephora (Rubiaceae). Am. J. Bot. 90: 153-157. Klein, A., Steffan-Dewenter, I., Buchori, D., & Tscharntke, T. 2002. Effects of Land-Use Intensity in Tropical Agroforestry Systems on Coffee Flower-Visiting and Trap- Nesting Bees and Wasps. Conservation Biology 16: 1003-1014. Klein, A., Vaissiere, B. E., Cane, J. H., Steffan-Dewenter, I., Cunningham, S. A., Kremen, C., & Tscharntke, T. 2007. Importance of pollinators in changing landscapes for world crops. Proceedings of the Royal Society of London B 274: 303-313. Kleinman, P. J. A., Pimentel, D., & Bryant, R. B. 1995. The ecological sustainability of slash-and-burn agriculture. Agriculture, Ecosystems & Environment 52: 235-249. Kowalewski, M., Serrano, G. E. A., Flessa, K. W., & Goodfriend, G. A. 2000. Dead delta's former productivity: Two trillion shells at the mouth of the Colorado River. Geology 28: 1059-1062. Kremen, C. 2005. Managing ecosystem services: what do we need to know about their ecology? Ecology Letters 8: 468-479. Kremen, C., & Chaplin-Kramer, R. 2007. Insects as providers of ecosystem services: crop pollination and pest control. In Stewart, A.J.A., New, T.R., & Lewis, O.T. (Eds.), Insect Conservation Biology: proceedings of the Royal Entomological Society's 23rd Symposium. Wallingford, UK: CABI Publishing (pp. 349-382). Kremen, C., Daily, G. C., Klein, A., & Scofield, D. in press. Inadequate assessment of the ecosystem service rationale for conservation: a reply to Ghazoul. Conservation Biology. Kremen, C., Williams, N. M., & Thorp, R. W. 2002. Crop pollination from native bees at risk from agricultural intensification. Proceedings of the National Academy of Sciences 99: 16812-16816. 187 Kremen, C., Williams, N. M., Aizen, M. A., Gemmill-Herren, B., LeBuhn, G., Minckley, R., Packer, L., Potts, S. G., Roulston, T., Steffan-Dewenter, I., Vazquez, D. P., Winfree, R., Adams, L., Crone, E. E., Greenleaf, S. S., et al. 2007. Pollination and other ecosystem services produced by mobile organisms: a conceptual framework for the effects of land-use change. Ecology Letters 10: 299-314. Kremen, C., Williams, N. M., Bugg, R. L., Fay, J. P., & Thorp, R. W. 2004. The area requirements of an ecosystem service: crop pollination by native bee communities in California. Ecology Letters 7: 1109-1119. Kruess, A., & Tscharntke, T. 1994. Habitat fragmentation, species loss, and biological control. Science 264: 1581-1584. Kunkel, C. M., Hallberg, R. W., & Oppenheimer, M. 2006. Coral reefs reduce tsunami impact in model simulations. Geophysical Research Letters 33: L23612. Landis, D. A., Wratten, S. D., & Gurr, G. M. 2000. Habitat management to conserve natural enemies of arthropod pests in agriculture. Annual review of entomology 45: 175-201. Larsen, T. H., Williams, N. M., & Kremen, C. 2005. Extinction order and altered community structure rapidly disrupt ecosystem functioning. Ecology Letters 8: 538- 547. Laurance, W. F. 2007. Environmental science: Forests and floods. Nature 449: 409-410. Lawton, R. O., Nair, U. S., Pielke, R. A., & Welch, R. M. 2001. Climatic impact of tropical lowland deforestation on nearby montane cloud forests. Science 294: 584-587. Lee, S., Ryu, J., & Kim, I. 2007. Landslide susceptibility analysis and its verification using likelihood ratio, logistic regression, and artificial neural network models: case study of Youngin, Korea. Landslides 4: 327-338. Lehner, B., & D6ll, P. 2004. Development and validation of a global database of lakes, reservoirs and wetlands. Journal of Hydrology 296: 1-22. Liang, X., Lettenmaier, D. P., Wood, E. F., & Burges, S. J. 1994. A simple hydrologically based model of land surface water and energy fluxes for general circulation models. Journal of Geophysical Research 99: 14415-14428. Lindsey, P. A., Alexander, R., Frank, L. G., Mathieson, A., & Romanach, S. S. 2006. Potential of trophy hunting to create incentives for wildlife conservation in Africa where alternative wildlife-based land uses may not be viable. Animal Conservation 9: 283-291. Lindsey, P., Roulet, P., & Romanach, S. 2007. Economic and conservation significance of the trophy hunting industry in sub-Saharan Africa. Biological Conservation 134: 455- 469. Losey, J. E., & Vaughan, M. 2006. The Economic Value of Ecological Services Provided by Insects. BioScience 56: 311-323. 188 MacArthur, R.H., & Wilson, E.O. (1967) The Theory of Island Biogeography. New Jersey: Princeton. Mader, P., Fliessbach, A., Dubois, D., Gunst, L., Fried, P., & Niggli, U. 2002. Soil fertility and biodiversity in organic farming. Science 296: 1694-1697. Malakoff, D. 1998. Coastal Ecology: Death by suffocation in the Gulf of Mexico. Science 281: 190-192. Mayaux P., Holmgren P., Achard F., Eva H., Stibig H.-J. & Branthomme A. 2005. Tropical forest cover change in the 1990s and options for future monitoring. Philosophical Transactions of the Royal Society of London B 360: 373-384. Mayaux, P., Holmgren, P., Achard, F., Eva, H., Stibig, H., & Branthomme, A. 2005. Tropical forest cover change in the 1990s and options for future monitoring. Philosophical Transactions of the Royal Society B: Biological Sciences 360. Mazda, Y., Magi, M., Ikeda, Y., Kurokawa, T., & Asano, T. 2006. Wave reduction in a mangrove forest dominated by Sonneratia sp. Wetlands Ecology and Management 14: 365-378. McArthur L. C, Boland J. W, Brown B, & Jones G. K. 2003. Investigation of the influence of seagrass on fisheries catch level using isotonic regression. Environmental Modeling and Assessment 8: 285-290. McCorquodale, S. M. 1997. Cultural contexts of recreational hunting and native subsistence and ceremonial hunting: their significance for wildlife management. Wildlife Society Bulletin 25: 568-573. McCullough, D. R. 1996. Spatially structured populations and harvest theory. The Journal of Wildlife Management 60: 1-9. McIntyre, P. B., Jones, L. E., Flecker, A. S., & Vanni, M. J. 2007. Fish extinctions alter nutrient recycling in tropical freshwaters. Proceedings of the National Academy of Sciences 104: 4461-4466. Mclvor, A. 2004. Freshwater mussels as biofilters. PhD Thesis. Retrieved from http://www.ourearth.co.uk/annamcivor/thesis.html MEA (Millennium Ecosystems Assessment). 2005. Ecosystems and Human Well-being: Synthesis Report. Island Press. Memmott, J., Craze, P. G., Waser, N. M., & Price, M. V. 2007. Global warming and the disruption of plant-pollinator interactions. Ecology Letters 10: 710-717. Merode, E., Homewood, K., & Cowlishaw, G. 2004. The value of bushmeat and other wild foods to rural households living in extreme poverty in Democratic Republic of Congo. Biological Conservation 118: 573-581. Michalski, F.. & Peres, C.A. 2007. Disturbance-mediated mammal persistence and abundance-area relationships in Amazonian forest fragments. Conservation Biology 21: 1626-1640. 189 Micheli, F., & Halpern, B. S. 2005. Low functional redundancy in coastal marine assemblages. Ecology Letters 8: 391-400. Millennium Ecosystem Assessment 2005. Ecosystems and Human Wellbeing: Biodiversity Synthesis. World Resources Institute, Washington DC. Milner-Gulland, E. J.. & Akgakaya, H. R. 2001. Sustainability indices for exploited populations. Trends in Ecology & Evolution 16: 686-692. Milner-Gulland, E. J., Bennett, E. L., Abernethy, K., Bakarr, M., Bodmer, R. E., Brashares, J. S., Cowlishaw, G., Elkan, P., Eves, H. E., Fa, J. E., Milner-Gulland, E. J., Peres, C., Roberts, C. M., Robinson, J. G., Rowcliffe, J. M., et al. 2003. Wild meat: the bigger picture. Trends in Ecology & Evolution 18: 351-357. Mitsch, W. J., Day, J. W., Gilliam, J. W., Groffman, P. M., Hey, D. L., Randall, G. W., & Wang, N. 2001. Reducing Nitrogen Loading to the Gulf of Mexico from the Mississippi River Basin: Strategies to Counter a Persistent Ecological Problem. BioScience 51: 373-388. Moberg, F., & Folke, C. 1999. Ecological goods and services of coral reef ecosystems. Ecological Economics 29: 215-233. Montag, J. M., Patterson, M. E. and Freimund, W. A. 2005. The wolf viewing experience in the Lamar Valley of Yellowstone National Park. Human Dimensions of Wildlife 10: 273-84. Moore, G. and Jennings, S. (eds.) 2000. Commercial Fishing: The Wider Ecological Impacts. London, UK: Blackwell Science. Moran, D. 1994. Contingent valuation and biodiversity: measuring the user surplus of Kenyan protected areas. Biodiversity & Conservation 3: 663-684. Morandin, L. A., & Winston, M. L. 2006. Pollinators provide economic incentive to preserve natural land in agroecosystems. Agriculture, Ecosystems & Environment 116: 289-292. Morandin, L. A., Winston, M. L., Abbott, V. A., & Franklin, M. T. 2007. Can pastureland increase wild bee abundance in agriculturally intense areas? Basic and Applied Ecology 8: 117-124. Morato, T, Cheung, W. W. L., Pitcher, T. J. 2004. Vulnerability of Seamount Fish to Fishing: Fuzzy analysis of life history attributes. In T. Morato & D. Pauly (Eds) Seamounts: Biodiversity and fisheries. Nancouver, Canada: Fisheries Centre, University of British Columbia (pp. 52-60). Morato, T. & Pauly, D. (eds). 2004. Seamounts: Biodiversity and fisheries. Fisheries Centre Research Reports 12: 1-78. Morato, T., Watson, R., Pitcher, T. and Pauly, D. 2006. Fishing down the deep. Fish and Fisheries 7: 24-34. 190 Moreira, F., Siqueira, J., & Brussaard, L. (Eds.). 2006. Soil Biodiversity in Amazonian and Other Brazilian Ecosystems. Wallingford, UK: CABI. Morrissey, M. M., Wieczorek, G. F., & Morgan, B. A. 2008. A comparative analysis of simulated and observed landslide locations triggered by Hurricane Camille in Nelson County, Virginia. Hydrological Processes 22: 524-531. Msangi, S., Sulser, T., Rosegrant, M., Valmonte-Santos, R., 2007. Global scenarios for biofuels: Impacts and implications for food security and water use. 10th Annual Conference on Global Economic Analysis, Purdue University, USA. Muchaal, P. K. & Ngandjui, G. 1999. Impact of village hunting on wildlife populations in the Western Dja Reserve, Cameroon. Conservation Biology 13: 385-396. Muir, J. 1898. The wild parks and forest reservations of the west. Atlantic monthly 81: 15. Mumby, P. J., Edwards, A. J., Ernesto Arias-Gonzalez, J., Lindeman, K. C., Blackwell, P. G., Gall, A., Gorezynska, M. I., Harborne, A. R., Pescod, C. L., Renken, H., C. C. Wabnitz, C., & Llewellyn, G. 2004. Mangroves enhance the biomass of coral reef fish communities in the Caribbean. Nature 427: 533-536. Mumby, P. J., Hastings, A.. & Edwards, H. J. 2007. Thresholds and the resilience of Caribbean coral reefs. Nature 450: 98-101. Myers, R. A., & Worm, B. 2003. Rapid worldwide depletion of predatory fish communities. Nature 423: 280-283. Myers, R. A., Baum, J. K., Shepherd, T. D., Powers, S. P., & Peterson, C. H. 2007. Cascading effects of the loss of apex predatory sharks from a coastal ocean. Science 315: 1846-1850. Nabhan, G. P., and S. L. Buchmann. 1997. Services provided by pollinators. Pages 133-150 in G. C. Daily, ed. Nature's services. Island Press, Washington, D.C. Naidoo R., Balmford A., Costanza R., Fisher B., Green R.E., Lehner B., Malcolm T.R. & Ricketts T.H. in press. Global Mapping of Ecosystem Services and Conservation Priorities. Proceedings of the National Academy of Sciences, USA. Naidoo, R., & Adamowicz, W. L. 2005. Economic benefits of biodiversity exceed costs of conservation at an African rainforest reserve. Proceedings of the National Academy of Sciences USA 102: 16712-16716. Naidoo, R., & Ricketts, T. H. 2006. Mapping the economic costs and benefits of conservation. PLoS Biology 4: e360 EP -. Naidoo, R., Balmford, A., Costanza, R., Fisher, B., Green, R. E.. Lehner, B., Malcolm, T. R., & Ricketts, T. H. in press. Global Mapping of Ecosystem Services and Conservation Priorities. Proceedings of the National Academy of Sciences. Naiman, R. J., & Decamps, H. 1997. The ecology of interfaces: riparian zones. Annual Review of Ecology and Systematics 28: 621-658. 19] Naiman, R. J., Melillo, J. M., & Hobbie, J. E. 1986. Ecosystem alteration of boreal forest streams by Beaver (Castor canadensis). Ecology 67: 1254-1269. Nakicenovic, N., Swart, R. (eds) 2000. IPCC Special Report on Emissions Scenarios. Cambridge University Press, Cambridge. Naughton-Treves, L., Mena, J. L., Treves, A., Alvarez, N., & Radeloff, V. C. 2003. Wildlife Survival Beyond Park Boundaries: the Impact of Slash-and-Burn Agriculture and Hunting on Mammals in Tambopata, Peru. Conservation Biology 17: 1106-1117. Navrud, S., & Mungatana, E. D. 1994. Environmental valuation in developing countries: The recreational value of wildlife viewing. Ecological Economics 11: 135-151. Naylor, R. & Ehrlich, P. 1997. Natural pest control services and agriculture. In: Daily, G. (ed.) Nature's Services: Societal Dependence on Natural Ecosystems. Island Press, Washington (pp. 151-174). Neigel, J. E. 2003. Species-Area Relationships and Marine Conservation. Ecological Applications 13: 138. Newton, K., Coté, I.M., Pilling, G.M., Jennings, S., Dulvy, N.K. 2007 Current and future sustainability of island coral reef fisheries. Current Biology 17:655-658. NIES (National Institute for Environmental Studies). 2008. The 13th AIM (Asia-Pacific Integrated Model) International Workshop. 16-18 February 2008, Ohyama Memorial Hall, National Institute for Environmental Studies, Tsukuba, Japan. Nijssen, B., Schnur, R., & Lettenmaier, D. P. 2001. Global Retrospective Estimation of Soil Moisture Using the Variable Infiltration Capacity Land Surface Model, 1980-93. Journal of Climate 14: 1790-1808 . Njiru, M., Kazungu, J., Ngugi, C. C., Gichuki, J., & Muhoozi, L. 2008. An overview of the current status of Lake Victoria fishery: Opportunities, challenges and management strategies. Lakes & Reservoirs: Research & Management 13: 1-12. NOAA (National Oceanic and Atmospheric Administration). 2008. NOAA Pacific Tsunami Warning Center. Retrieved April 15, 2008, from http://www.prh.noaa.gov/pr/ptwe/ Novaro, A. J., Funes, M., & Walker, R. S. 2005. An empirical test of source-sink dynamics induced by hunting. Journal of Applied Ecology 42: 910-920. Novaro, A. J., Redford, K. H., & Bodmer, R. E. 2000. Effect of hunting in source-sink systems in the Neotropics. Conservation Biology 14: 713-721. NRC (National Research Council). 2007. Status of Pollinators in North America. Washington D.C., USA: National Academies Press. Ntiamoa-Baidu, Y. 1997. Wildlife and food security in Africa. Rome, Italy: Food and Agriculture Organization of the United Nations [online] http://www.fao.org/docrep/w7540e/w7540e00.htm#Contents. 192 Ohl-Schacherer, J., Shepard, G. H., Kaplan, H., Peres, C. A., Levi, T., & Yu, D. W. 2007. The sustainability of subsistence hunting by Matsigenka native communities in Manu National Park, Peru. Conservation Biology 21: 1174-85. Orr, J. C., Fabry, V. J., Aumont, O., Bopp, L., Doney, S. C., Feely, R. A., Gnanadesikan, A., Gruber, N., Ishida, A., Joos, F., Key, R. M., Lindsay, K., Maier-Reimer, E., Matear, R., Monfray, P., et al. 2005. Anthropogenic ocean acidification over the twenty-first century and its impact on calcifying organisms. Nature 437: 681-686. Ostman, O., Ekbom, B., & Bengtsson, J. 2003. Yield increase attributable to aphid predation by ground-living polyphagous natural enemies in spring barley in Sweden. Ecological Economics 45: 149-158. Palomares, F., Gaona, P., Ferreras, P., & Delibes, M. 1995. Positive effects on game species of top predators by controlling smaller predator populations: an example with lynx, mongooses, and rabbits. Conservation Biology 9: 295-305. Palumbi, S. R. 2001. Humans as the world's greatest evolutionary force. Science 293: 1786- 1790. Pandolfi, J. M., Bradbury, R. H., Sala, E., Hughes, T. P., Bjorndal, K., Cooke, R. G., McArdle, D., McClenachan, L., Newman, M. J. H., Paredes, G., Warner, R. R., & Jackson, J. B. C. 2003. Global trajectories of the long-term decline of coral reef ecosystems. Science 301: 955-958. Pandolfi, J. M., Jackson, J. B. C., Baron, N., Bradbury, R. H., Guzman, H. M., Hughes, T. P., Kappel, C. V., Micheli, F., Ogden, J. C., Possingham, H. P., & Sala, E. 2005. Are U.S. Coral reefs on the slippery slope to slime? Science 307: 1725-1726. Parry, M. L., Canziani, O. F., Palutikof, J. P., van der Linden, P. J., Hanson, C. E. (Eds.) 2007. Climate Change 2007 - Impacts, Adaptation and Vulnerability: Working Group II contribution to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge. Pauly, D. and Watson, R. 2005 Background and Interpretation of the ‘Marine Trophic Index’ as a measure of biodiversity. Phil. Trans. Soc. B, 360, 415-423. Pauly, D., Alder, J. Bakun, A., Heileman, S., Kock, K., Mace, P., Perrin, W., Stergiou, K., Sumaila, U. R., Vierros, M., Freire, K., Sadovy, Y., Christensen, V., Kaschner, K., Palomares, M., et al. 2005 Marine Fisheries Systems. In Millennium Ecosystem Assessment. Ecosystems and Human Well-being: Current States and Trends. Washington D.C., USA: Island Press (pp. 745-794). Pauly, D., Alder, J., Bennett, E., Christensen, V., Tyedmers, P., & Watson, R. 2003. The future for fisheries. Science 302: 1359-1361. Pauly, D., Christensen, V., Dalsgaard, J., Froese, R., & Torres, F. 1998. Fishing down marine food webs. Science 279: 860-863. Pawson, M. G., Tingley, D., & Glenn, H. 2004. “Sport Fisheries” (or Marine Recreational Fisheries) in the EU. Brussels, Belgium: The European Commission Directorate- 193 General for Fisheries. Online: http://www.eu.nl/fisheries/publications/studies/sportfishing 2006.pdf Peres, C. A. & Dolman, P. M. 2000. Density compensation in neotropical primate communities: evidence from hunted and unhunted Amazonian forests of varying productivity. Oecologia 122: 175-189. Peres, C. A. 1999. Effects of subsistence hunting and forest types on Amazonian primate communities. In J. G. Fleagle, C. Janson & K. E. Reed (Eds) Primate Communities. Cambridge, UK: Cambridge University Press (pp. 268-283). Peres, C. A. 2000. Effects of subsistence hunting on vertebrate community structure in Amazonian forests. Conservation Biology 14: 240-253. Peres, C. A. 2001. Synergistic effects of subsistence hunting and habitat fragmentation on Amazonian forest vertebrates. Conservation Biology 15: 1490-1505. Peres, C. A., & Palacios, E. 2007. Basin-wide effects of game harvest on vertebrate population densities in Amazonian forests: implications for animal-mediated seed dispersal. Biotropica 39: 304-315. Pergams, O. R. W., & Zaradic, P. A. 2008. Evidence for a fundamental and pervasive shift away from nature-based recreation. Proceedings of the National Academy of Sciences USA Petley, D., Hearn, G., Hart, A., Rosser, N., Dunning, S., Oven, K., & Mitchell, W. 2007. Trends in landslide occurrence in Nepal. Natural Hazards 43: 23-44. Petley, D.N., Dunning, S.A. & Rosser, N.J. 2005. The analysis of global landslide risk through the creation of a database of worldwide landslide fatalities. In Hungr, O., Fell, R., Counture, R. & Ebergardt, E. (Eds.) Landslide Risk Management. Amsterdam: Balkema (pp. 367-374). Petrosillo, I., Zurlini, G., Corliano, M., Zaccarelli, N., & Dadamo, M. 2007. Tourist perception of recreational environment and management in a marine protected area. Landscape and Urban Planning 79: 29-37. Pimentel, D. 2008. Conservation biological control. Biological Control 45: 171. Pimentel, D., Berger, B., Filiberto, D., Newton, M., Wolfe, B., Karabinakis, E., Clark, S.. Poon, E., Abbett, E., & Nandagopal, S. 2004. Water resources: agricultural and environmental issues. BioScience 54: 909-918. Pinet, J. 1995. The hunter in Europe. FACE: Federation of Associations for Hunting and Conservation of the EU. Pitcher, T. J., Watson, R., Forrest, R., Valtysson, H. P., & Guenette, S. 2002. Estimating illegal and unreported catches from marine ecosystems: a basis for change. Fish and Fisheries 3: 317-339. 194 Pitcher, T., & Cochrane, K. 2002. The use of ecosystem models to investigate multispecies management strategies for capture fisheries. Fisheries Centre Research Reports 10: 1- 115. Plaganyi, E. E. 2007. Models for an ecosystem approach to fisheries. Rome: Food and Agriculture Organization of the United Nations. Postel, S. 1999. Pillar of Sand: Can the Irrigation Miracle Last? New York, USA: W.W Norton & Co. Postel, S. L., & Thompson, B. H. 2005. Watershed protection: Capturing the benefits of nature's water supply services. Natural Resources Forum 29: 98-108. Potting, J., Bakkes, J. (eds) 2004. The GEO-3 Scenarios 2002-2032 Quantification and Analysis of Environmental Impacts. RIVM, Bilthoven, Netherlands. Potts, S. G., Vulliamy, B., Roberts, S., O'Toole, C., Dafni, A., Ne'eman, G., & Willmer, P. 2005. Role of nesting resources in organising diverse bee communities in a Mediterranean landscape. Ecological Entomology 30: 78-85. Pregitzer, K. S., & Euskirchen, E. S. 2004. Carbon cycling and storage in world forests: biome patterns related to forest age. Global Change Biology 10: 2052-2077. Priess, J. A., Mimler, M., Klein, A. M., Schwarze, S., Tscharntke, T., & Steffan-Dewenter, I. 2007. Linking deforestation scenarios to pollination services and economic returns in coffee agroforestry systems. Ecological applications 17: 407-17. Pusey, B. J., & Arthington, A. H. 2003. Importance of the riparian zone to the conservation and management of freshwater fish: a review. Marine and Freshwater Research 54: 1-16. Rabalais, N. N., Turner, E. R., & Wiseman, W. R. 2002. Gulf of Mexico hypoxia, a.k.a. “The Dead Zone”. Annual Review of Ecology and Systematics 33: 235-263. Ramankutty, N., & Foley, J. A. 1999. Estimating historical changes in global land cover: Croplands from 1700 to 1992. Global Biogeochemical Cycles 13: 997-1027. Ramankutty, N., Evan, A. T., Monfreda, C., & Foley, J. A. 2008. Farming the planet: 1. Geographic distribution of global agricultural lands in the year 2000. Global Biogeochemical Cycles 22: GB1003. Redford, K. H. 1992. The empty forest. BioScience 42: 412-422. Richards, A. J. 2001. Does Low Biodiversity Resulting from Modern Agricultural Practice Affect Crop Pollination and Yield? Annals of Botany 88: 165-172. Ricketts, T. H. 2004. Tropical Forest Fragments Enhance Pollinator Activity in Nearby Coffee Crops. Conservation Biology 18: 1262-1271. Ricketts, T. H., Daily, G. C., Ehrlich, P. R., & Michener, C. D. 2004. Economic value of tropical forest to coffee production. Proceedings of the National Academy of Sciences 101: 12579-12582. 195 Ricketts, T. H., Regetz, J., Steffan-Dewenter, I., Cunningham, S. A., Kremen, C., Bogdanski, A., Gemmill-Herren, B., Greenleaf, S. S., Klein, A. M., Mayfield, M. M., Morandin, L. A., Ochieng’, A., & Viana, B. F. 2008. Landscape effects on crop pollination services: are there general patterns? Ecology Letters 11: 499-515. Ricketts, T. H., Williams, N. M., & Mayfield, M. M. 2006. Connectivity and ecosystem services: crop pollination in agricultural landscapes. In K. R. Crooks & M. Sanjayan (Eds.), Connectivity Conservation. Cambridge, UK: Cambridge University Press (pp. 255-289). Rizkalla, C., Blanco-Silva, F., & Gruver, S. 2007. Modeling the impact of ebola and bushmeat hunting on western lowland gorillas. EcoHealth 4: 151-155. Roberts, C. M., Andelman, S., Branch, G., Bustamante, R. H., Carlos Castilla, J., Dugan, J., Halpern, B. S., Lafferty, K. D., Leslie, H., Lubchenco, J., McArdle, D., Possingham, H. P., Ruckelshaus, M., & Warner, R. R. 2003. Ecological criteria for evaluating candidate sites for marine reserves. Ecological Applications 13: 199. Roberts, C. M., Bohnsack, J. A., Gell, F., Hawkins, J. P., & Goodridge, R. 2001. Effects of Marine Reserves on Adjacent Fisheries. Science 294: 1920-1923. Robinson, J. G., & Bennett, E. L. 2004. Having your wildlife and eating it too: an analysis of hunting sustainability across tropical ecosystems. Animal Conservation 7: 397-408. Robinson, J. G., & Redford, K. H. 1986b. Intrinsic rate of natural increase in Neotropical forest mammals: relationship to phylogeny and diet. Oecologia 68: 516-520. Robinson, J. G., & Redford. K. H. 1986a. Body Size, Diet, and Population-Density of Neotropical Forest Mammals. American Naturalist 128:665-680. Robinson, J. G., Redford, K. H., & Bennett, E. L. 1999. Wildlife harvest in logged tropical forests. Science 284: 595-596. Rogers, C. S. 1990. Response of coral reefs and reef organisms to sedimentation. Mar. Ecol. Prog. Ser. 62, 185-202. Roldan, A., & Simonetti, J. 2001. Plant-mammal interactions in tropical Bolivian forests with different hunting efforts. Cons 15: 617-623. Ronnback, P. 1999. The ecological basis for economic value of seafood production supported by mangrove ecosystems. Ecological Economics 29: 235-252. Rosegrant, M. W., Agcaoili-Sombilla, M., Perez, N. D., 1995 Global Food Projections to 2020: Implications for Investment. Diane Publishing. Rosell, F., Bozser, O., Collen, P. and Parker, H. 2005 Ecological impact of beavers Castor fiber and Castor Canadensis and their ability to modify ecosystems. Mammal Rev. 35: 248-276. Roubik, D.W. 2002. Feral African bees augment neotropical coffee yield. In Pollinating Bees: The Conservation Link Between Agriculture and Nature (eds P. Kevan & V.L. Imperatriz-Fonseca). Brasilia, Brazil: Ministry of Environment (pp. 255-266). 196 Roughgarden, J., & Smith, F. 1996. Why fisheries collapse and what to do about it. Proceedings of the National Academy of Sciences USA 93: 5078-5083. Rounsevell, M. D. A., Reginster, I., Araijo, M. B., Carter, T. R., Dendoncker, N., Ewert, F., House, J. I., Kankaanpaa, S., Leemans, R., Metzger, M. J., May 2006. A coherent set of future land use change scenarios for Europe. Agriculture, Ecosystems & Environment 114: 57-68. Rounsevell, M., Reginster, I., Araujo, M., Carter, T., Dendoncker, N., Ewert, F., House, J., Kankaanpaa, S., Leemans, R., Metzger, M., Schmit, C., Smith, P., & Tuck, G. 2006. A coherent set of future land use change scenarios for Europe. Agriculture, Ecosystems & Environment 114: 57-68. Rowcliffe, J. M., Cowlishaw, G., & Long, J. 2003. A model of human hunting impacts in multi-prey communities. Journal of Applied Ecology 40: 872-889. Rowcliffe, J. M., Milner-Gulland, E., & Cowlishaw, G. 2005. Do bushmeat consumers have other fish to fry? Trends in Ecology & Evolution 20: 274-276. Rowell, K., Flessa, K. W., Dettman, D. L., & Roman, M. 2005. The importance of Colorado River flow to nursery habitats of the Gulf corvina (Cynoscion othonopterus). Canadian Journal of Fisheries and Aquatic Sciences 62: 2874-2885. RSPB (Royal Society for the Protection of Birds). 2006. The RSPB: Help us halt illegal hunting: France. Online: http://www.rspb.org.uk/supporting/campaigns/ illegalhunting/france.asp (April 26, 2008) Runte, A. 1987. National Parks: the American experience. Lincoln & London: University of Nebraska Press. Rusuwa, B., Maruyama, A. and Yuma, M. 2006 Deterioration of cichlid habitat by increased sedimentation in the rocky littoral zone of Lake Malawi. Ichthyol Res 53: 431-434. Ryder, R. A. (1965) A Method for Estimating the Potential Fish Production of North- temperate Lakes. Transactions of the American Fisheries Society 94: 214-218. Sandker, M., Suwarno, A., Campbell, B. M., 2007. Will forests remain in the face of oil palm expansion? simulating change in Malinau, Indonesia. Ecology and Society 12 (2), 37+ Sandker, M., Suwarno, A., Campbell, B. M., 2007. Will forests remain in the face of oil palm expansion? simulating change in Malinau, Indonesia. Ecology and Society 12 (2), 37+ Sathaye, J., Chan, P., Dale, L., Makundi, W., Andrasko, K., 2003. A summary note estimating global forestry GHG mitigation potential and costs: A dynamic partial equilibrium approach. Economic and Environmental Modeling Workshop, 19-20 January, 2004. Habitat Center, New Delhi. Saunders, M. A., & Lea, A. S. 2008. Large contribution of sea surface warming to recent increase in Atlantic hurricane activity. Nature 451: 557-560. 197 SAUP (Sea Around Us Project). 2008. Sea Around Us Project. Online: http://www.seaaroundus.org/ (accessed 3 March 2008). SCBD. 2003. Interlinkages between biological diversity and climate change: Advice on the integration of biodiversity considerations into the implementation of the United Nations Framework Convention on Climate Change and its Kyoto protocol. Montreal, Canada: Secretariat of the Convention on Biological. Scheffer, M., Carpenter, S., & Young, B. D. 2005. Cascading effects of overfishing marine systems. Trends in Ecology & Evolution 20: 579-581. Scheffer, M., Carpenter, S., Foley, J. A., Folke, C., & Walker, B. 2001. Catastrophic shifts in ecosystems. Nature 413: 591-596. Scherr, S. J., & McNeely, J. A. 2008. Biodiversity conservation and agricultural sustainability: towards a new paradigm of 'ecoagriculture' landscapes. Philosophical Transactions of the Royal Society of London. Series B, Biological sciences 363: 477- 94. Schjonning, P., Elmholt, S.. & Christensen, B. T. (Eds.). 2004. Managing Soil Quality: Challenges in Modern Agriculture. CABI. Schroéter, D., Acosta-Michlik, L., Arnell, A., Aratjo, M., Badeck, F., Bakker, M., Bondeau, A., Bugmann, H., Carter, T., Vega-Leinert, A. D. L., Erhard, M., Espifieira, G., Ewert, F., Fritsch, U., Friedlingstein, P., et al. 2004. Advanced Terrestrial Ecosystem Analysis and Modelling - Final Report 2004 (Section 5 and 6 and Annex | to 6). Potsdam, Germany: Potsdam Institute for Climate Impact Research. Online: http://www.pik-potsdam.de/ateam/ateam_final_report_sections 5 to 6.pdf Schuttenberg, H. Z. 2001. Coral Bleaching; Causes, Consequences and Response: Selected Papers presented at the 9th International Coral Reef Symposium on “Coral Bleaching: Assessing and Linking Ecological and Socioeconomic Impacts, Future Trends and Mitigation Planning”. Rhode Island, USA: Coastal Resources Centre, University of Rhode Island. SEI, no date. Scenarios for sustainability. Stockholm Environment Institute US Center. Online: http://www.scenariosforsustainability .org/. (accessed 18 April 2008). Seitzinger, S. P., Harrison, J. A.. Dumont, E., Beusen, A. H. W., Bouwman, A. F., 2005. Sources and delivery of carbon, nitrogen, and phosphorus to the coastal zone: An overview of global nutrient export from watersheds (news) models and _ their application. Global Biogeochemical Cycles 19: GB4S01+ Shennan, C. 2008. Biotic interactions, ecological knowledge and agriculture. Philosophical transactions of the Royal Society of London. Series B, Biological sciences 363: 717- 39. Sheppard, C., Dixon, D. J., Gourlay, M., Sheppard, A., & Payet, R. 2005. Coral mortality increases wave energy reaching shores protected by reef flats: Examples from the Seychelles. Estuarine, Coastal and Shelf Science 64: 223-234. 198 Shrestha RK, Stein TV, Clark J. 2007. Valuing nature-based recreation in public natural areas of the Apalachicola River region, Florida. Journal of Environmental Management 85: 977-985. Sidle, R. C., Ziegler, A. D., Negishi, J. N., Nik, A. R., Siew, R., & Turkelboom, F. 2006. Erosion processes in steep terrain--Truths, myths, and uncertainties related to forest management in Southeast Asia. Forest Ecology and Management 224: 199-225. Sierra, R., Rodriguez, F., & Losos, E. 1999. Forest resource use change during early market integration in tropical rain forests: the Huaorani of upper Amazonia. Ecological Economics 30: 107-119. Silveira, R., & Thorbjarnarson, J. B. 1999. Conservation implications of commercial hunting of black and spectacled caiman in the Mamiraua Sustainable Development Reserve, Brazil. Biological Conservation 88: 103-109. Simpson, R. L., Good, R. E., Walker, R., & Frasco, B. R. 1983. The role of Delaware River freshwater tidal wetlands in the retention of nutrients and heavy metals. J Environ Qual 12: 41-48. Sinclair, A. R. E., & Krebs, C. J. 2002. Complex numerical responses to top-down and bottom-up processes in vertebrate populations. Philosophical Transactions of the Royal Society B: Biological Sciences 357. Smith, K. G. & Darwall, W. R.T. 2006. The Status and Distribution of Freshwater Fish Endemic to the Mediterranean Basin. \UCN, Gland, Switzerland and Cambridge, UK. Smith, R. S., Shiel, R. S., Bardgett, R. D., Millward, D., Corkhill, P., Evans, P., Quirk, H., Hobbs, P. J., & Kometa, S. T. 2008. Long-term change in vegetation and soil microbial communities during the phased restoration of traditional meadow grassland. Journal of Applied Ecology 45: 670-679. Soares-Filho, B. S., Nepstad, D. C., Curran, L. M., Cerqueira, G. C., Garcia, R. A., Ramos, C. A., Voll, E., Mcdonald, A., Lefebvre, P., Schlesinger, P., 2006. Modelling conservation in the Amazon basin. Nature 440 (7083), 520-523. Soares-Filho, B. S., Nepstad, D. C., Curran, L. M., Cerqueira, G. C., Garcia, R. A., Ramos, C. A., Voll, E., Mcdonald, A., Lefebvre, P., Schlesinger, P., 2006. Modelling conservation in the Amazon basin. Nature 440 (7083), 520-523. Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M., Miller, H.L. (Eds.) 2007. Climate Change 2007: The Physical Science Basis. Contribution of Working Group I Contribution to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge. Spalding, M. D., Ravilious, C., & Green, E. P. (Eds.). 2001. World Atlas of Coral Reefs. Berkley, CA, USA: University of California Press. Retrieved April 17, 2008, from http://www.unep-wemce.org/marine/seagrassatlas/index.htm Spalding, M., Blasco, F., & Field, C. (Eds.). 1997. World Mangrove Atlas. Okinawa, Japan: International Society of Mangrove Ecosystems. 199 Spotila, J. R., Reina, R. D., Steyermark, A. C., Plotkin, P. T., Paladino, F. V. 2000 Pacific leatherback turtles face extinction. Nature 405:529-530. Srinivasan, U. T., Carey, S. P., Hallstein, E., Higgins, P. A. T., Kerr, A. C., Koteen, L. E., Smith, A. B., Watson, R., Harte, J., & Norgaard, R. B. 2008. The debt of nations and the distribution of ecological impacts from human activities. Proceedings of the National Academy of Sciences 105: 1768-1773. Stachowicz J., Bruno, J. F. & Duffy, J. E. 2007 Understanding the effects of marine biodiversity on communities and ecosystems. Annual Review of Ecology, Evolution, and Systematics 38: 739-766. Stachowicz, J. J., Whitlatch, R. B., & Osman, R. W. 1999. Species diversity and invasion resistance in a marine ecosystem. Science 286: 1577-1579. Steffan-Dewenter, I. 2003. Importance of Habitat Area and Landscape Context for Species Richness of Bees and Wasps in Fragmented Orchard Meadows. Conservation Biology 17: 1036-1044. Steffan-Dewenter, I., Munzenberg, U., Burger, C., Thies, C., & Tscharntke, T. 2002. Scale- Dependent Effects of Landscape Context on Three Pollinator Guilds. Ecology 83: 1421-1432. Stein, C. J., & Colditz, G. A. 2004. The Epidemic of Obesity. J Clin Endocrinol Metab 89: 2522-2525. Steinfeld, H., Gerber, P., Wassenaar, T., Castel, V., Rosales, M., & de Haan, C. 2006. Livestock’s long shadow: Environmental issues and options. Livestock, Environment and Development (LEAD) Initiative of the Food and Agriculture Organization (FAO). Stern, N. 2007. The Economics of Climate Change: The Stern Review. Cambridge University Press. Stevenson, S. 2007. The European Fishing Industry. Brussels, Belgium: European Parliamant. Online: http://www.scottishtorymeps.org.uk/archive/stevenson/issues/reports/NewEUFish.pdf Stocks, K. 2004. Seamountsonline: An online resource for data on the biodiversity of seamounts. In T. Morato & D. Pauly (Eds) Seamounts: Biodiversity and fisheries. Vancouver, Canada: Fisheries Centre, University of British Columbia (pp. 13-16). Stoner, K. E., Riba-Hernandez, P., Vulinec, K., & Lambert, J. E. 2007. The role of mammals in creating and modifying seedshadows in tropical forests and some possible consequences of their elimination. Biotropica 39: 316-327. Straub, C. S., & Snyder, W. E. 2006. Species identity dominates the relationship between predator biodiversity and herbivore suppression. Ecology 87: 277-282 . Stuart, S. N., Chanson, J. S., Cox, N. A., Young, B. E., Rodrigues, A. S. L., Fischman, D. L., & Waller, R. W. 2004. Status and Trends of Amphibian Declines and Extinctions Worldwide. Science 306: 1783-1786. 200 Sundaravadivel, M., & Vigneswaran, S. 2001. Constructed wetlands for wastewater treatment. Critical Reviews in Environmental Science and Technology 31: 351-409. Sunderland, K., & Samu, F. 2000. Effects of agricultural diversification on the abundance, distribution, and pest control potential of spiders: a review. Entomologia Experimentalis et Applicata 95: \-13. Swift, M. J., Izac, A. N., & van Noordwijk, M. 2004. Biodiversity and ecosystem services in agricultural landscapes--are we asking the right questions? Agriculture, Ecosystems & Environment 104: 113-134. Synolakis, C. E., & Kong, L. 2006. Runup measurements of the December 2004 Indian Ocean tsunami. Earthquake Spectra 22: S67-S91. Takano, T., Nakamura, K., & Watanabe, M. 2002. Urban residential environments and senior citizens’ longevity in megacity areas: the importance of walkable green spaces. J Epidemiol Community Health 56: 913-918. Tanaka, N., Sasaki, Y., Mowjood, M., Jinadasa, K., & Homchuen, S. 2007. Coastal vegetation structures and their functions in tsunami protection: experience of the recent Indian Ocean tsunami. Landscape and Ecological Engineering 3: 33-45. Ten Brink, B., Alkemade, R., Bakkenes, M., Eickhout, B., de Heer, M., Kram, T., Manders, T., van Oorschot, M., Smout, F., Clement, J., van Vuuren, D., Westhoek, H., Miles, L., Lysenko, I., Fish, L., et al. 2006. Cross-roads of Planet Earth. Exploring means to meet the 2010 biodiversity target. Montreal, Canada: Secretariat of the Convention on Biological Diversity. Terborgh, J., Lopez, L., Nunez, P., Rao, M., Shahabuddin, G., Orihuela, G., Riveros, M., Ascanio, R., Adler, G. H., Lambert, T. D., & Balbas, L. 2001. Ecological meltdown in predator-free forest fragments. Science 294: 1923-1926. The World Bank. 2005. Going, Going, Gone: The Illegal Trade of Wildlife in East and Southeast Asia. Washington D.C., USA: The World Bank. Thies, C., Roschewitz, I., & Tscharntke, T. 2005. The landscape context of cereal aphid— parasitoid interactions. Proceedings of the Royal Society B: Biological Sciences 272. Tilman, D., Reich, P. B., & Knops, J. M. H. 2006. Biodiversity and ecosystem stability in a decade-long grassland experiment. Nature 441: 629-632. Tockner, K., & Stanford, J. A. 2002. Riverine flood plains: present state and future trends. Environmental Conservation 29: 308-330. Tscharntke, T., Bommarco, R., Clough, Y., Crist, T. O., Kleijn, D., Rand, T. A., Tylianakis, J. M., Nouhuys, S. V., & Vidal, S. 2007. Conservation biological control and enemy diversity on a landscape scale. Biological Control 43: 294-309. Tscharntke, T., Gathmann, A., & Steffan-Dewenter, I. 1998. Bioindication using trap- nesting bees and wasps and their natural enemies: community structure and interactions. The Journal of Applied Ecology 35: 708-719. 201 Tscharntke, T., Klein, A. M., Kruess, A., Steffan-Dewenter, I., & Thies, C. 2005. Landscape perspectives on agricultural intensification and biodiversity - ecosystem service management. Ecology Letters 8: 857-874. Tscharntke, T., Rand, T., & Bianchi, F. 2005. The landscape context of trophic interactions: insect spillover across the crop - noncrop interface. Ann. Zool. Fennici 42: 421-432. Tylianakis, J. M., Didham, R. K., & Wratten, S. D. 2004. Improved fitness of aphid parasitoids receiving resource subsidies. Ecology 85: 658-666. Tylianakis, J. M., Tscharntke, T., & Lewis, O. T. 2007. Habitat modification alters the structure of tropical host-parasitoid food webs. Nature 445: 202-205. Ulrich, R. S., Simons, R. F., Losito, B. D., Fiorito, E., Miles, M. A., & Zelson, M. 1991. Stress recovery during exposure to natural and urban environments. Journal of Environmental Psychology 11: 201-230. UNDP (United Nations Development Programme). 2006. Human Development Report 2006. Beyond scarcity: Power, poverty and the global water crisis. New York, NY, USA: United Nations Development Programme. UNEP (United Nations Environment Programme). 2005. After the Tsunami: Rapid Environmental Assessment. Nairobi, Kenya: United Nations Environment Programme. UNESCO-WWAP (United Nations Educational, Scientific and Cultural Organization — World Water Assessment Programme). 2006. The 2nd UN World Water Development Report: 'Water, a shared responsibility’. New York, NY, USA: United Nations Educational, Scientific and Cultural Organization — World Water Assessment Programme. USFW (U.S. Fish & Wildlife Service). 2007. 2006 Nationai Survey of Fishing, Hunting, and Wildlife-Associated Recreation - State Overview (Preliminary Findings). U.S. Fish & Wildlife Service. USFWS 2007. 2006 National Survey of Fishing, Hunting, and Wildlife-Associated Recreation: State Overview. US Fish and Wildlife Service, Washington D.C. Uyarra, MC, Cote, IM, Gill, JA, Tinch, RRT, Viner, D, & Watkinson, AR. 2005. Island- specific preferences of tourists for environmental features: Implications of climate change for tourism-dependent states. Environmental Conservation 32: 11-19. Valiela, I., Bowen, J. L., & York, J. K. 2001. Mangrove Forests: One of the World's Threatened Major Tropical Environments. BioScience 51: 807-815. van Dijk, P. P., Stuart, B. L., & Rhodin, A. G. 2000. Asian Turtle Trade: Proceedings of a Workshop on Conservation and Trade of Freshwater Turtles and Tortoises in Asia. Lunenburg, MA, USA: Chelonian Research Foundation. Van Voorhees, D., & Pritchard, E. S. 2007. Fisheries of the United States 2006. Silver Spring, MD, USA: National Marine Fisheries Service, Fisheries Statistics Division. 202 van Wilgen, B. W., Cowling, R. M., & Burgers, C. J. 1996. Valuation of ecosystem services: a case study from South African fynbos ecosystems. BioScience 46: 184- 189. Vazquez, D. P., Morris, W. F., & Jordano, P. 2005. Interaction frequency as a surrogate for the total effect of animal mutualists on plants. Ecology Letters 8: 1088-1094. Veerman, J. L., Barendregt, J. J., van Beeck, E. F., Seidell, J. C., & Mackenbach, J. P. 2007. Stemming the Obesity Epidemic: A Tantalizing Prospect. Obesity 15: 2365- 2370. Verburg, P. H., Soepboer, W., Veldkamp, A., Limpiada, R., Espaldon, V., Mastura, S. S., 2002. Modeling the spatial dynamics of regional land use: The clue-s model. Environmental Management 30 (3), 391-405. VGrésmarty, C. J., Lévéque, C., Revenga, C., Bos, R., Caudill, C., Chilton, J., Douglas, E. M., Meybeck, M., Prager, D., Balvanera, P., Barker, S., Maas, M., Nilsson, C., Oki, T., Reidy, C. A., et al. 2005. Fresh Water. In Millennium Ecosystem Assessment. Ecosystems and Human Well-being: Current States and Trends. Washington D.C., USA: Island Press (pp. 165-207). V6résmarty, C., Federer, C., & Schloss, A. 1998. Potential evaporation functions compared on US watersheds: Possible implications for global-scale water balance and terrestrial ecosystem modeling. Journal of Hydrology 207: 147-169. Wade, DJ, Mwasaga, BC & Eagles, PFJ. 2001. A history and market analysis of tourism in Tanzania. Tourism Management 22: 93-101. Waite, T. 2007. Revisiting evidence for sustainability of bushmeat hunting in West Africa. Environmental Management 40: 476-480. Walpole, M. J., & Leader-Williams, N. 2002. Tourism and flagship species in conservation. Biodiversity and Conservation 11: 543-547. Walpole, M.J. & Thouless, C.R. 2005. Increasing the value of wildlife through non- consumptive use? Deconstructing the myths of ecotourism and community-based tourism in the tropics. In People & Wildlife: Conflict or Coexistence (Eds R.Woodroffe, s. Thirgood & A.Rabinowitz). Walpole, M.J. 1997. Dragon tourism in Komodo National Park, Indonesia: its contribution to conservation and local development. PhD thesis, University of Kent, Canterbury, UK. Walsh, P. D., Abernethy, K. A., Bermejo, M., Beyers, R., De Wachter, P., Akou, M. E., Huijbregts, B., Mambounga, D. I., Toham, A. K., Kilbourn, A. M., Lahm, S. A., Latour, S., Maisels, F., Mbina, C., Mihindou, Y., et al. 2003. Catastrophic ape decline in western equatorial Africa. Nature 422: 611-614. Watson, R., Coles, R. G., & Long, W. J. L. 1993. Simulation estimates of annual yield and landed value for commercial penaid prawns from a tropical seagrass habitat, Northern Queensland, Australia. Aust. J. Mar. Freshwater Res. 44: 211-219. 203 Watson, R., Morato, T. 2004. Exploitation Patterns in Seamount Fisheries: A preliminary analysis. In T. Morato & D. Pauly (Eds) Seamounts: Biodiversity and fisheries. Vancouver, Canada: Fisheries Centre, University of British Columbia (pp. 61-66). Welcomme, R. L. 1999 A review of a model for qualitative evaluation of exploitation levels in multi-species fisheries. Fisheries Management and Ecology 6:1—-19. Welcomme, R. L., & Halls, A. 2004. Dependence of tropical river fisheries on flow. In R. L. Welcomme & T. Petr (Eds.), International Symposium on the Management of Large Rivers for Fisheries. Phnom Pehn, Cambodia 11-14 Feb 2005. Bangkok, Thailand: Food and Agriculture Organization of the United Nations (FAO) and Mekong River Commission (MRC). Wells, M.P. 1993. Neglect of biological riches: the economics of nature tourism in Nepal. Biodiversity & Conservation 2: 445-464. Wells, S., Ravilious, C., & Corcoran, E. 2006. In the front line: Shoreline protection and other ecosystem services from mangroves and coral reefs . Cambridge, UK: UNEP- World Conservation Monitoring Centre. Westerkamp, C., & Gottsberger, G. 2000. Diversity Pays in Crop Pollination. Crop Science 40: 1209-1222. Whitman, K., Starfield, A. M., Quadling, H. S., & Packer, C. 2004. Sustainable trophy hunting of African lions. Nature 428: 175-178. Wilby, A., & Thomas, M. B. 2002. Natural enemy diversity and pest control: patterns of pest emergence with agricultural intensification. Ecology Letters 5: 353-360. Wilby, A., Villareal, S. C., Lan, L. P., Heong, K. L., & Thomas, M. B. 2005. Functional benefits of predator species diversity depend on prey identity. Ecological Entomology 30: 497-501. Wilkie, D. S. & Lee, R. J. 2004. Hunting in agroforestry systems and landscapes: conservation implications in West-Central Africa and Southeast Asia. In G. Schroth, G. A. B. Da Fonseca, C. A. Harvey, H. L. Vasconceles, & A.-M. N. Izac (Eds.). Agroforestry and biodiversity conservation in tropical landscapes. Washington DC: Island Press. (pp 346-370) Wilkie, D. S. 2005. Hunting in agroforestry systems and landscapes: conservation implications in West-Central Africa and Southeast Asia. In G. Schroth, G. da Fonseca, C. Harvey, C. Gascon, H. L. Vasconcelos, & A. M. N. Izac (Eds.), Agroforestry and biodiversity conservation in tropical landscapes. Washington D.C., USA: Island Press (pp. 346-370). Wilkie, D. S., & Carpenter, J. 1999a. The potential role of safari hunting as a source of revenue for protected areas in the Congo Basin. Oryx 33: 340-345. Wilkie, D. S., & Carpenter, J. F. 1999b. Bushmeat hunting in the Congo Basin: an assessment of impacts and options for mitigation. Biodiversity and Conservation 8: 927-955. 204 Wilkinson C. (Ed.) 2002. Status of coral reefs of the world: 2002. Townsville, Queensland: Global Coral Reef Monitoring Network, Australian Institute of Marine Sciences. Wilkinson, C. (Ed.) 2000, Status of Coral Reefs of the World: 2000 Townsville, Queensland, Australia: Australian Institute of Marine Science. Willer, H., & Yussefi, M. (Eds.). 2008. The World of Organic Agriculture - Statistics and Emerging Trends 2008. Frick, Switzerland: International Federation of Organic Agriculture Movements (IFOAM), DE-Bonn and Research Institute of Organic Agriculture, FiBL. Online: from http://orgprints.org/13123. Willer, H., Yussefi-Menzler, M. & Sorensen, N. (Eds). 2008. The World of Organic Agriculture - Statistics and Emerging Trends 2008. International Federation of Organic Agriculture Movements (IFOAM), DE-Bonn and Research Institute of Organic Agriculture, FiBL, CH-Frick. Williams-Guillen, K., Perfecto, I., & Vandermeer, J. 2008. Bats limit insects in a neotropical agroforestry system. Science 320: 70. Wilson, E.O. 1984. Biophilia. Harvard University Press, Cambridge, MA. Winfree, R., Griswold, T., & Kremen, C. 2007a. Effect of Human Disturbance on Bee Communities in a Forested Ecosystem. Conservation Biology 21: 213-223. Winfree, R., Williams, N. M., Dushoff, J.. & Kremen, C. 2007b. Native bees provide insurance against ongoing honey bee losses. Ecology Letters 10: 1105-1113. Wint, W., & Robinson, T. 2006. Gridded Livestock of the World. Rome, Italy: Food and Agriculture Organization of the United Nations. With, K. A., Pavuk, D. M., Worchuck, J. L., Oates, R. K., & Fisher, J. L. 2002. Threshold effects of landscape structure on biological control in agroecosystems. Ecological Applications 12: 52-65 . Wood, S. Ehui, S., Alder, J., Benin, S., Cassman, K. G., Cooper, H. D., Johns, T., Gaskell, J., Grainger, R., Kadungure, S., Otte, J., Rola, A., Watson, R., Wijkstrom, U., Devendra, C., Kanbar, N. et al. 2005 Food. In: Millennium Ecosystem Assessment. Ecosystems and Human Well-being: Current States and Trends. Washington D.C., USA: World Resources Institute (pp. 209-241). Wood, S., Ehui, S., Alder, J., Benin, S., Cassman, K. G., Cooper, D. H., Johns, T., Gaskell, J., Grainger, R., Kadungure, S., Otte, J., Rola, A., Watson, R., Wijkstrom, U., Devendra, C., et al. 2005. Food. In Millennium Ecosystem Assessment. Ecosystems and Human Well-being: Current States and Trends. Washington D.C., USA: Island Press (pp. 209-241). Wood, S., Sebastian, K., & Scherr, S. J. Pilot analysis of global ecosystems: Agroecosystems. Washington D.C., USA: International Food Policy Research Institute (IFPRI) and World Resources Institute. Woolley, H., & Rose, S. 2004. The Value of Public Space. London, UK: CABE Space. 205 World Health Organisation. 2006. Obesity and overweight factsheet. Retrieved from http://www.who.int/mediacentre/factsheets/fs3 | 1/en/index.html. Worm, B., Barbier, E. B., Beaumont, N., Duffy, J. E., Folke, C., Halpern, B., Jackson, J. B. C., Lotze, H. K., Micheli, F., Palumbi, S. R., Sala, E., Selkoe, K. A., Stachowicz, J. J., & Watson, R. 2006. Impacts of Biodiversity Loss on Ocean Ecosystem Services. Science 314: 787-790. Wright, J. P., & Flecker, A. S. 2004. Deforesting the riverscape: the effects of wood on fish diversity in a Venezuelan piedmont stream. Biological Conservation 120: 439-447. WTTC 2007. The global travel and tourism summit. World Travel and Tourism Council, London. UK. Yamano, Miyajima, & Koike. 2000. Importance of foraminifera for the formation and maintenance of a coral sand cay: Green Island, Australia. Coral Reefs 19: 51-58. Young, I. R., & Hardy, T. A. 1993. Measurement and modelling of tropical cyclone waves in the Great Barrier Reef. Coral Reefs 12: 85-95. Zavaleta, E. S. 2000. The economic value of controlling an invasive shrub. Ambio 29: 462- 467. Zedler, J. B., & Kercher, S. 2005. Wetland resources: status, trends, ecosystem services, and restorability. Annual Review of Environment and Resources 30: 39-74. Zhang, W., Ricketts, T. H., Kremen, C., Carney, K., & Swinton, S. M. 2007. Ecosystem services and dis-services to agriculture. Ecological Economics 64: 253-260. Zhu, Y., Chen, H., Fan, J., Wang, Y., Li, Y., Chen, J., Fan, J., Yang, S., Hu, L., Leung, H., Mew, T. W., Teng, P. S., Wang, Z., & Mundt, C. C. 2000. Genetic diversity and disease control in rice. Nature 406: 718-722. 206 Annex 1 — Cost and benefits of agriculture conversion If is not straightforward to decide how to deal with systems (such as agricultural fields) that despite being heavily modified by human activities still hold substantial biodiversity and produce ecosystem services. For example, it is not clear if the economic value of wheat production should be considered a benefit of the ecosystem service ‘food production’ or an opportunity cost of not converting a forest to a wheat field. This raised considerable discussion in each of the three consultations we had on the conceptual framework. This discussion turned out to be extremely productive, as it became clear that it does not matter how outputs from agricultural fields are classified — what is fundamental is to ensure the adequacy of the ‘states’ being compared (Section 2). Indeed, as long as these states are equal in everything else but the implementation or not of a set of conservation actions, the end economic consequences are the same irrespective of whether wheat production (for example) is treated as beneficial service enhanced by habitat conversion, or as an opportunity cost of not converting. We illustrate this with a simplified example of a parcel of land with two possible states: a forest plot and a wheat field. Taken separately, each of these states has a set of running benefits and costs (lists of costs and benefits have been simplified for illustration purposes). The benefits of the forest plot are that it can produce a sustainable timber harvest (FB), stores carbon (FB2) and it provides habitat for wild bees that pollinate nearby crops (FB3). The costs of running the forest are expenses in forest management such as patrolling (FC,) and paying compensation to nearby farmers for the damage caused by forest animals (FC2). The benefits of the wheat field are wheat production (WBj), and some carbon storage (WBz), while the costs are those related to crop management, such as ploughing the field (WC,) and paying to clean the pollution from agricultural runoff to nearby rivers (WC). Note that ‘timber production’ and ‘wheat production’ are here simply as economic benefits of particular types of land use, irrespective of whether they are considered ecosystem services or not. In the same way, “damage to nearby crops’ and ‘pollution runoff are treated here as costs rather than ‘ecosystem disservices’. Suppose the question being asked is: what are the net consequences of converting forest to wheat? As described in the general framework (Section 2), a comparison between states can be done by investigating differences in benefits and in costs in both states. When converting from forest to wheat (Figure A1), the differences in benefits are: e The loss of the benefit of timber production (- FB1); e The added benefit of wheat production (+ WB1); e The difference in carbon storage (+ WB2 - FB2); and e The loss of the benefit of pollination of nearby crops (- FB3). In addition, there is a one-off benefit of conversion, the windfall of timber harvesting when removing the forest (+ TH). 207 The net difference in benefits (A B) is the sum of each of these parcels. rte hdd « fost timber production = ~ FB, « changed management costs = + WC, ~ FC, added wheat production = + WB, * avoided damage to nearby crops = — FC; . + changed carbon storage = + WE,-FB, + added pollution runott tc naarby nvers = + WC, ¢ ost pollination of nearby crops = - FB, Differences in benafits (€}- Differences in costs (€): isa tel: One-off cost of conversion {€}: * windfall of one-off timber harvest = + TH * costs of physical conversion = + PC Mot diffarances in boncfits (€): Net differences in costs (€}: 4B =—FB, + WB, + WB,- FB, - FB, + TH « AC=WC, -FC,—FC, + WC, + PC he nnnurenneansenenennanmesean ae A€= AB - AC =— FB, + WB, + WE2- FB, - FB; + TH- (WC; —FC, FC, + WC; * PC) = WE,+ WB, + FC, + FC; + TH — WC,-WC; - FB,-FB:—FB; + PC ee ee added wheat avelded = 1-off timber added lost forest 1-céf physical bereits foreatecats harvest wheat costs benefits converaion | advantages of conversion Gaadvantages cf conversion IF: AE <0 IF: AE >0 nat differences In bonofis < nat differences Im coats net differences im benofits > nat Giforences In costs sdvetlages of commrsion < dikmelwarlages of Conversion adhraMagia of converion > Gadvanbagea of coneralon net € loss from conversion s* net € gain from eS economic recommendation: not convert economic recommendation: convert Figure Al: Evaluating the net consequences of converting a forest plot to a wheat field. When converting from forest to wheat, the differences in costs are: e The difference in the costs of managing wheat instead of forest (+ WC1 - FC1); 208 e The avoided costs of compensating for damage to nearby crops (- FC2); and e The added costs of cleaning the pollution from agricultural runoff (+ WC2). There is also a one-off cost of physically converting the land to make it suitable for wheat production (e.g., removing tree roots; + PC). The net difference in costs (A C) is the sum of each of these parcels. As per the general framework (Figure 2, Section 2), the net consequences of conversion (A €) can be assessed by comparing the net differences in benefits (A B) with the net differences in costs (A C). If the former are larger than the latter, there is a net economic gain (A € > 0) from conversion, hence it makes economic sense to convert. The equation for the difference between net benefits and net costs can be rearranged to indicate as positive terms all the advantages of conversion (the added benefits from wheat, the avoided forest costs, and the one-off windfall in timber) and as negative terms all the disadvantages of conversion (the added costs from wheat, the lost benefits of forest, and the one-off costs of land conversion). If the advantages are larger than the disadvantages, it makes economic sense to convert (Figure A1). Note that ‘wheat production’ was simply an added economic benefit of converting forest to wheat, irrespective of whether it is classified as an ecosystem service or not. Suppose the question being asked is: what are the net consequences of not converting forest to wheat? Importantly, the question is not “what are the net consequences of retaining the forest” (these would simply be a continuation of state A), but the net consequences of actions put in place to conserve the forest plot from conversion to a specific alternative form of land use. Only by framing the question in this way it is possible to account for the opportunity costs of conserving the forest. The differences in benefits when not converting to a wheat field are perfectly symmetrical to the differences in the benefits of conversion, and so are the differences in costs (Figure A2). When computing the net consequences of no conversion, all parcels that were previously the advantages of conversion (Figure Al) are now the disadvantages of no conversion (Figure A2), and vice-versa. The end result is exactly the same: the conditions that would result in the recommendation not to convert identical whichever way the question is put (Figure Al cf Figure A2). In the latter figure, the value of ‘wheat production’ simply becomes an opportunity cost of maintaining the forest. yas State A: Forest plot _ Differences in benefits (€}: * retained timber production =+ FB, + changed managamant costs = — WC, + FC, * prevented wheal production = — We, * maintained damage to nearby crops = + FC; + changed carbon storage = — WB, * FB; = avoided pollution runoff to nearby rivers = — WC, * retained pollination of nearby craps = + FBS * prevented windfall of timbar harvest =-— TH * avolded costs of physical conversion =-PC Net differences in benefits (€): Net differences in costs (€): AB = FB, —WE,— WE, + FB, + FB,—-TH AC = —WC, + FC, +FC,-Wwec, -Pc RASS OO RRR SERRE SOR ARORESRORR ERE R ADEE ERODE Eee nennne ene e ADRES ORER REAR RRROn sR annensenennoonsnnennl Peeeeret rrr reer e rt eeee errr 4 } Prevented one-off benatit ot conversion {€): (€): : H F A€ = AB — AC © FB, - WB, — WB, + FB, + FB;— TH —{- WC, + FC, * FC;—WC2-PC} = FB, +FB;+ FB, + WC, + Wo; PC - FC,-FC, - WE,-WB, - TH te ed _——— prevented prevented 1-off fornat benefits whestcasta T-of physical forest costs whest benefits winctall of conversion timber harvest nD | adwartages of nat converting disadvantages of not converting IF: A€ > 0 IF: A€<0 net differences in Senate > nat differences i coma net differences tm benefits < met differences In costs advantages of not converting > disadvantages of not convening 3¢ advantages of not converting < disadvantages of not convarting net € gain from not converting galt net € loss from not converting eo economic recommendation: not convert economic recommendation: convert Figure A2. Evaluating the net consequences of not converting a forest plot to a wheat field. As demonstrated, it is irrelevant which parcels are considered benefits from ecosystem services or not. The relevant information is whether they are benefits or costs of particular states. Accordingly, it does not matter what classification of ecosystem services is being used. Indeed, this same framework could have been used to assess the net economic consequences 210 of converting (or not converting) a forest to a car park, or traditional extensive wheat production to intensive wheat production. As discussed in Section 2, the fundamental step in ensuring that these comparisons address the remit of the Review is to ensure that the states being compared contrast two specific situations: one where conservation action to prevent wild nature loss has been put in place, and another where it has not. In the previous example of forest plot and wheat field, the comparison is relevant if it is deemed that conserving the forest is the action required for preventing biodiversity loss. The net economic consequences of conversion (Figure Al) can therefore be interpreted as the economic consequences of failing to protect biodiversity, while, conversely, the net economic consequences of no conversion (Figure A2) are the economic consequences of biodiversity loss. Annex 2 — Framework development and consultation We had three opportunities to present and discuss the framework with different audiences: - Seminar “Conceptualising ecosystem services and human well-being”, 15 January, Department of Geography, University of Cambridge. We presented a talk to an audience dominated by environmental economists and geographers and solicited additional comments by email. - Presentation “Framework for a Stern-like review of the economics of biodiversity loss”, 16 January, Department of Zoology, University of Cambridge. We presented a talk in one of our departmental lunchtime seminars. There were about 20 attendees, mainly biologists. - Zoology workshop: the framework was discussed in a small (8 people) workshop of key experts. A detailed account of these discussions is presented in the Inception Report to this project. The systematically raised concern was “the agriculture issue” — to include or not agricultural food production as an ecosystem service?). In these discussions we found what we believe is the solution for that problem (Section 2; Annex 1). Several other interesting issues were raised and discussed. Overall, the framework has proved remarkably robust. The discussions have also been very useful in making us think more carefully about the details of how the framework is applied (e.g., how to consider interactions and trade-offs between services). 212 Annex 3 - Practical strategy followed in the thematic reviews This Scoping the Science project required reviewing very quickly a wide literature, most of which out of the immediate areas of expertise of the members of the team. We therefore had two limiting factors: expertise and time. To make sure we access the relevant expertise, we had to draw on the support of experts for each theme, who know the literature well and have a broad perspective of the state of knowledge in their respective fields. Ideally, we would have wanted a team of experts in each theme producing the respective reports. This would have required a substantially longer and better-funded project than the current one. What was possible to obtain in the current project was more or less detailed guidance from a diversity of experts. Following such guidance required time, to being able to pursue the relevant references and to compile the overall information into a coherent report. The best way in which we were able to obtain researchers-time in such a short project was by hiring five interns. They were: James Beresford, Kelly Flower, Antares Hernandez, Hannah Peck, and James Waters. We were fortunate to hire five bright young biologists, but being very recent graduates they had only limited research experience. It would have been impossible to hire more experienced researchers in such short notice and given the limited funds available. This meant that their outputs had to be substantially edited before being sent back to the experts. Given time limitations, we were unable to fully develop all the themes considered. We considered that it was better to invest in producing fewer, but high-quality, reviews than to present lower quality reports for all themes. We therefore present eight fully developed thematic reviews, presenting only brief overviews for the remaining nine themes. The strategy adopted for those themes that were fully developed consisted of the following steps: Questionnaire A questionnaire (tailored to each task) was sent to a set of experts to try to obtain key pointers (to references, resources, other experts) as well as to capture the experts’ experience and opinion across a range of questions. First draft An intern reviewed the literature and the information provided by experts, and produced a first draft. Second draft Given the limited experience of the interns, the first drafts produced required considerable editing and double-checking of the information. This was done by Ana Rodrigues. A second draft was then produced. Review The second draft was sent back to the experts. Final report Comments from the experts are incorporated into a final report, that becomes a chapter in the final report of the Scoping the Science project. The contribution of experts is acknowledged in the following way: e Authors: experts who have contributed with substantial information and which reviewed and approved the final report. 213 e Contributors: experts who provided initial information but were unable to review the report, or experts that considered that this was a more suitable description of their contribution. e Reviewers: experts who reviewed the report but were not listed as authors. We gave experts the option to reproduce their comments at the end of each section if substantive comments were provided that we could not incorporate (e.g., given time limitations). Hence, being listed as a reviewed did not have to be considered an endorsement of the text. e Acknowledgements: experts who contributed with references and/or contacts but not with specific information, and who did not see the final text. For some the themes for which it was not possible to develop a full review, we present only a very brief overview of what we believe the state of the knowledge is, based on our own (limited) knowledge, as preliminary guidance for Phase 2. The overall level of expert support was extraordinary, given how busy these experts inevitably are. People could only afford to take the time because they understand the importance of the Review on the Economics of Biodiversity Loss. However, the extent to which we feel that we have succeeded in representing the state of knowledge in each field was very variable. In some cases (such as pollination) we feel that we have been able to tap into the knowledge and advice of the key people in the field, that they understood the purpose of this exercise, and that they took the time to provide substantial input. In other fields (e.g. wild harvested fibers) we feel that we were unable to obtain advice from key people and that therefore we may only present a partial view of the state of knowledge. Table Al indicates the level of information for each theme. For those that were fully developed, we also indicate the level of expert support. This is an assessment of the extent to which we feel that we have tapped significantly in the existing overall expertise in the field, and for which we are more confident that the review results are representative of the overall state of knowledge. Table Al Team (intern + head), level of information collected and level of expert support in each thematic review. Level of information Expert support JJ Waters + Ana Fully developed review Very good Rodrigues Wild crop pollination Biological control in JJ Waters + Ana Brief overview crops Rodrigues Genetic diversity crops | Hannah Peck + Ana Brief overview and livestock Rodrigues Soil quality for crop Antares Hernandez + | Brief overview production Ana Rodrigues Livestock JJ Waters + Ana Brief overview Rodrigues Marine fisheries Kelly Flower + Ana Fully developed review Rodrigues Inland fisheries Kelly Flower + Ana Fully developed review Rodrigues Wild animal products JJ Waters + Ana Fully developed review Rodrigues Fresh water provision Hannah Peck + Ana Fully developed review and regulation Rodrigues Timber from natural Antares Hernandez + | Not developed forests Ana Rodrigues Wild medicinal plants | Antares Hernandez + | Not developed Ana Rodrigues Nature-related outdoor | James Beresford + Fully developed review activities Matt Walpole Regulation of natural James Beresford + Not developed hazards Ana Rodrigues Very good Reasonable Antares Hernandez + Ana Rodrigues One-off use benefits Not developed Non-use benefits Ana Rodrigues Not developed Global climate Hannah Peck + Ana Brief overview regulation Rodrigues Unknown Ana Rodrigues Not developed benefits/processes 215 Annex 4 - Literature review “Economics of Biodiversity Loss” The research has been realized on the following search engines: e Ecological Economics e Nature e Scopus e Science e Science direct e Google scholar (succinctly) A combination of key words was used during the research for each search engine: e Biodiversity + Economic Loss e Biodiversity + Economic Cost e Biodiversity + Economic Valuation From these key words, the articles including all the terms in the text and published between 2005-2008 (included) have been selected. A total of 132 documents have been gathered, comprising: e 2005: 29 articles e 2006: 36 articles e 2007: 46 articles e 2008: 13 articles e Inpress: 8 articles The selected articles come from different sources the main ones being: e Agriculture, Ecosystems and Environment (5) e Agricultural and Resource Economics Review (2) e Biological Conservation (10) e Coastal Engineering (2) e Conservation Biology (6) 216 e Ecological economics, (21) e Ecological Modeling (3) e Environmental Resource Economics (4) e Environmental Modeling and Software (2) e Environmental Science and Policy (3) e Forest Ecology and Management (3) e Journal of Sustainable Forestry (5) e Land-Use Policy (2) e Marine Policy (4) e Marine pollution bulletin (3) e Nature (2) e PLOS Biology (2) e Science (14) The selected articles have been divided into two groups: e the articles about topics that concerns general issues (Table A2) e the articles related to case-studies in specific areas (Table A3) Authors Linda Scholten, Emmanuelle Noirtin & Rudolf de Groot (Environmental Systems Analysis group, Wageningen University) 817 86b-L8h OT oa solmouoga ]09180]09, Li 9007 6S JOA s [2918 0/00q uoreAsiasuoo AWISIOAIporg oF ABOjOOS puw soruouoda “y Suonoisueyy JO sojor oy} Burssassy ‘ommyeu jo Ayloswos oynjosqe pue aanepay pue Ye “Joqey “D Wyo Ss ‘Ioupes une g TC “paoy pur “AWD ‘OMed “A'O ‘eddey‘joeq “pO “Apouuoy “Wd ‘smog “y'T JoutwiD “sg ‘uemsy “Ss “Aysb[Od “7a yup “p ‘woaewug “A “Pysuejoyy “d's ‘soyoeH “Wg “URUTTTTIS “MA WON “aa Yorqueg sonje A pue suorjouny ECESIGE 8002 6TE TPA EES [BorZojooy swOUT[UON YIM JUoWOSeURY poseg-wo}sXsooq ][e}SeOD We Jouye A pue “Sy “uBA Ppeasivep “pd ‘Toyos “Wl “uedvfueg “qf “Zonsupoy €1Z-ZI1Z 6007 LOE TOA aoualog 2318], OLOT SAWSIOAIC [RoIBojorg UO UOUDAUOD ay], “HL ‘snaxory “M Psosped “TL ‘sdutyd “d “Surpoyy ; 5 : c : “q ‘xneAepy “Y “ARP “DH “oR “AL “ofoao'] SN SSWRIT[I\-Japeoy “Dp ‘uowory “qa “Apouuay “fp SuostuBy “Cy “Alosain “gy “usaIH “y] “uoynd “N ‘Aojpnq “yw ‘uosqoq “d “auBID “WT “A190 “q ‘isdoop “g ‘us) yuug “y ‘unuudg “Vy ‘psoywyeg * CAL Sones] ( 5, ae aA 81-0 9007 ee FOIA Aja100g ul Adojouyoa uonoayjar sdauonnoesd vy :Ayyiqeureysng wv ‘uojajddy c e-68E te bes Senet SISA]OU Fr AAUI]UO, ISROUTR MOUs or woe ee Z) ‘d ‘nssny pue *g ‘tsoysiog “Vy ‘loojuy B86E-68E $002 €9 JOA Se IMON Jeunjdo uv ur sormeuAp vo1sojoo9 puv o1wOUOdS UdAaMyaq UOHORIOU] (Oech nesnell poe SM sete se ava a epee # p-¢ Sonss] = = sorweudp jesnjeu Jo uoearasaid pue woysAs [Bo130]099 ay} ‘anbaagy] “pq “sa[Mouy “[-7 ‘ejeqemey “OW “Joussay “PV ‘“uo\suy “q ‘uoospng afPnes “H ‘uuvwidiog “sy ysuyor “y ‘ppomeay “WY “J9[syooq “A ‘PIozieM PUR D THO" Isyor “PW “ISYooI ‘ye 19 -Y “oz A ‘uideyS “fp ‘ouorsey “q ‘ueuLy, “gs ‘zed a ‘ue, pue yy ‘urdeyS yemg “yp ‘quordsey “g “zeIq “L ‘suyor pur ‘f ‘ossyjeqed “s “9p storg ‘d's “IIe “yy ezueysoD “L Opa pus yl YS “xy “Aydingy “fp “uaueyy “N “AaquBHE “JN “OSH Sutuueyd oarasor DAI}OATJO-]SOO JOF aN]BA oUIS JO ainsvow Jeorjowid VW :jS00 JUoUD0R|dOYy “y ‘uauR[Ioyy puke ‘Py “ezoqeD SYOHLIV LOOT-8Te PLe-S9e PLS-OLS PT9-179 6OILI-8ILI Sd9vd LOOT LOOT S007 LOOT ssoud uy L007 ssaud ut ssaud ut aLva 81 TOA STE TOA “p-¢ sonss] JOA 60E TOA €-Z sans] 19 TOA WANNTOA aoualog aoualg LYSA10 +] ajqouinjsng fo jpu.inor anualag UOIjNJOAG PUY Asojooy ul SCNAYL SOMOUOIT [0I1BO}OIA aoualog SOMMOUOIT ppaidojoo. Y SONUOUOIT 102130]007 @GONNOS 1 POUSTADY UOTBTTO[xAIDAG JO sorMOUoDA PHOA\ SursueySD ve ur Adojooy puelasur yy { Wed] OM PIp JeYA\ ‘SOdIAIOg WaysAsOoW JOJ sjoyswUT SuTsI9 WA, 28) pue’y Jo saouanbasuo,D jeqo[H des Aqypiqeureysns ouy puryy SUTYRUW-UOISINOP JOJ S9dIAJOS W9}SAsOoa SuUTATISse[O pue SUIUTZOC] AjISIOATPOIG SAIASUOD OF sus ed yoorC] MATADI [BOLO VY :UOIBAIASUOD pur SSO] AVSIOAIPOIG JO S[APOW SIWIOUODSS UT SIOJLOIPUL PUR saLIOdY] [BoISo[ooq suondo pue soduayjeyo JO MOIAaI [eqo]s & :sease pojoojoid Jo Suroueuly a[qvureysng puejesZ Moan Wo sojduexe yIM ‘sjoafoid paseqg-oIs Puv soqis JO on[eA UOHBAIOSUOD OY} SuTZeUISo JOJ youoidde odvospur] VW “MY Woy} pur “y ‘sedwoy “Oy ‘uoywH “| ‘ueuyoH pue "] “uosT|IH ‘Sa ‘Anquan “Sd JopAug pue “Ny “Aynyuewey OD] “sonuag “wf ‘ze “O “‘eparyuoyy “PO Yueyony “Wea “‘pwemop, “, “Aemol[OH HL “PISMONSH “STH “§qqID"O'D “APEC LIN 90D “Sy ‘uldeyo° ys “ayuadiep “DH ‘ueuog “) ‘piopeg “qb ‘ousy"y ‘sowed “yr ‘Ao}o4 “V ‘ope pur “y ‘uauayog “7 “UIGOY “Lf “PIO “EL “MEMON “al “SIN “q ‘preuoqoryy “qq “edewuopury “yy “uyosuroyy “gq eUPID “sg Howry “y ‘uoyay oy yporueq “Ed ‘asoy “A “UaAIS “CV “Buruueyy “yp OYst d “Sulployy pue “SY “Jowiny, “yy ‘ezueysoH “Eg “oYysy “YW “SSI pue ‘f'q “O1w.u0,7 NT Of “Yop wea ysiog pue ay yurddy “JT ‘sewoyy pue fp ‘doysig “yy ‘uopowy ‘{PUOUDAG “DoW pur qO-V ‘Hassny “yp ‘puowsq SYOHLIV CCT ber-Stp 8007 61E 1OA aouatog BUlOg-[19M\ A[quureysng Joy ASojouyos |, pur doust0g ‘df “USUP]OH] WW WBZ puw ‘q Jopeayeg-sey[t A, “7 ‘undsy sauoy, op “Vy ‘za1q-owloy “DH ‘souor “yw ‘pudy “fp ‘odue5-osouieny “4 ‘turmespus, “q doyfing “py ‘eusylig “YY ‘Apu “Nn “Andopumndieyy -2019g = “"S — “Waqqed DHE ‘uUassifawoD “W ‘sapeyo “OW ‘awowoYy-zaiog “Yl ‘zourpeyy -onsaeyy “OD “Joulag “q ‘zarq-osep “qd ‘sais9RD “A “opiqey “Vy ‘pavdog “y's ‘puvg “fd “UosTIA\ “df ‘own “WW ‘opiqeD “s “zeiq “y ‘uosdwoyy, “fp UIMOTTeEL, “DH “eyp-jeuosjuoyy “Hf “uosspopy uolprtasuoD jor1sojo1g {ATISIOAIp puv[ssess OAvs OF JSOO 4 [[IM YON MOP] SUOTJBJOPISUOD O1OUDs puL [eoIToTOoa Sururquioo ‘O'" “ue pueyio7 Soouc Pel peol Berea ‘uoroajoid AISIOAIPOIG JO Suryuvsl o1wMOUCSS sjuRWUZ}IOA\ SUIpUdIXy| pue “LOL Uop ueA YysIog “JO “Jap uea oplopy SOMMOUOI TT StII-SOIL [PIUAULUOAIAUT JUSWUOTAUT SY} pue sorMOUoDA amBJOM\ [eIodwopajUy [Z Jo\deyo 1D eo yooqpup 4 = 7 , R We ‘s8dug pue “yx ‘Teg “fp nA “TD ‘ueMpay 09L-9SL 6IE TOA aouatog SBSH eR CU ACME) eee cermemmims) “sie GSN Oa ‘MUTE Z anss] SP TOA Jonuoo [eorsojorq “LN ‘semoyy, 602-002 uoHeAIasuod Joy syeyqey J9}[9YS JO solwouocs pue Aousloly pue “vy “Aapeg “Wir “PuelorH “sro “syyUD J0.4NUOD JHII180]01g 1B 19 N WeL “O-O “IPHuyos “g “eneg “DH “uIpEYY Rg “Vy “eULIEg “Vp ‘upysnopyy “yy ‘uossenyEyy “PY ‘ouryeyy “D/O “ulpuny “p ‘deoaqry “] ‘sedwoy “3 19ya][° “fp ‘ydosop “J, ‘saaoin “g ‘wlomD “JN ‘sumer “Cd ‘sauinbg “] “Aumospry “Yy “WogttH “OY “YoyLIH p onssy “TOK Ao1jOg AULD POM sursueyo v ur souoysty SuruolIsog f QANTOA @OUNOs SYOHLNV OISI-eIst Lit TOA “p-€ sonss] €-Z Sanssy ¢-] sonss) 9II-ZOI val : O9T-TEl Z9-PS L007 Gieaae: s “ 6b% TOA Lr6l-trol S007 OIE OA sere S007 SL IOA ] onssy LZI-LIT JOA Sd9vd aLva FIANNTOA QIUALISOL aoualg SAMO 4] ajqouinssng fo Jpuinor SOMMOUOIT [DMBO]OIA JuawMasnuvnyy pun Asojooq jsaioy juaudojaaaq pun juauuowaug fo poudnor Juawuaspunypy pun Asoj09q salon] anualag ABOUOY 10974 UOlNDALASUO) joasojo1g @OuNOs €CT syoasut Aq paprAoid saorAdjas [vo1BO[O99 JO anjBA O1MIOUODD SY], sulo}sAg Jeinjen pue ueunyzy paydnoy jo Ayxajduop SOOIAING Wa}sAsooy JO sjayeyy SurAoidwy $1800 AjISIOAIPOIG UleLOOUN YIM MOIS puv UOTsIDAUOD pur] [eUTdO ydaouoo Jeorsojo09 UB UO UONNGLUOS soWOUODS jsAIOJ W :PoJOPISUOdAI s]SAOJ PAXIPY suiajsXg uoljonpoig A}IPpoMWOD-O15 Po}o9]9g JO sjordwy] orMOUODSOINDOG puke AjISIOAIPOIg, sodeospury [einqjno aurdyje ayy yo Alojs ay) :Asnpur pue soualos Jo aouanyyur oy} Jopun juowoseuRU jsosoy [BUOT PEI, uonensanbag uoqieD [eolsojorg YIM UogIeD JOF J9jeA\ Super], o8pajMouy yuo JO snsuasuoo y :Suruoyouny wojsXsooa uo AYISIOAIPOIG JO sjooTyA ajeos [euorsal e uo adevospury jeanypnouse ue UI SUOT}OB JUOWDSeUBUE UOTBAIDSUOD JO ssaUdATOaJJO-JSOO SureNnjeaq ATLIL ‘(ZW ‘ueysne, pure “q'f ‘Aaso7 ; “AM ‘JojAe], pue “P'S “Joprlouyog “TD ‘uewmpay “jy “1oyousrolg “7 “Suvdng “q ‘WoNsQ “f ‘oousyoqn’] “1 ‘qery “gq ‘uempeag “NCy ‘Ted “a “uesopy “5 “OMOd “IN “MHogiy “y's “YoyuodseD “YL “Zier “y “avy “A “BywAT puw “Vy “tuBAoH “WV ‘playuayory “dg “stung “y Hpeddny “yy ‘spoyuoqyory ‘¢ ‘Apoadp pure -q’y ‘xnolo7] WW ‘Soopy pur *D “owury “g ‘wumg “] ‘oyouy W soudyyeg pure “], sunquajayay, “L, pooy “ff sssjsso-y “gq ‘uueyor ‘O'd “Avunyy pur “weg YEON “O'd Oy ame “WM “Ade “AV “YOO “TC “NeLeg “as “Aoy “y yessiay “Oy “ASeqqor “gy “uOsyour e198 “g ‘Msneyouy y oper “py ‘Pema “g'q curdeyo “pq “sadooy] “y ‘qeddag pue vy “1odweyzjoy, SHYOHLIV 00S-06F ILOI-esol ST9-PT9 LOb-98E 807-Sol 9007 $007 ssoud ut L007 Z anssy TOA TT-OT sonss] (és TOA 8-L sonssy TOA Z-[ sonssy TOA S00T- 9007 [ onssy GANNTOA ABojolg UolpALasuoD BUlsaaulaug ]DjsvoD HNO S°Td BAnJON Aouogd puv aoualag JOJUaUlUO.AUT ABO]OIg Uo1jA.LasUuoD uOIDALLAasUuOD JOIBOjO1g AUISiaalpolg UONALASUO uoljonyo 4 SOMUOUOIT poaisoj0oq Ayssaalpolg Malay aounosay [ANNI SOIMOUOIT pun GOUNOS VCC Sodvospue Ty Jeuonisuezy, ul uonBaiosuod Jo sysop Ayunyoddo surjapoyy UBISAP BLIO}LIO [BOITO]OOS Sursy] ast] SULIEU JOF syEyGeY [BIOIIPe se saINjonss asuayap [B]SBOO Po}sos0-MO'T sso'T Ajisuaaiporg sjoiparg Aqjenbouy orwiouooy dine UO JNO SuIJag $]89J0J poumo Ajayeaud soy suesdoid sanuaour uoNeArasuoo AjISIOAIpOrg, sroquinyy jo Suruvsyy oy} suoneAsosuoD AysioaIporg Jo uonenjeA o1mouodg WOT}BAIASUOD AYISIDAIPOIq soy Avd 0) ssousuypim oy} purysq soayour o1mouosa-uou oy], MOIA JO JUIO O1WOUONT UL WOLF APISIOAIPOIg JO UOBAIASUOD [RqO[H spur] Suizess Jo juowoseuew pourejsns Joy [Borsojoos = puke = olwoUdDa ~=— BuI]IOUDDDY S}oI[yuoo dnoip Suryio,y spuary, pue uoNIpuoD oy) Jo sSuipuly :spuay pue oes UoLIND :Sulaq-]J9M UeWINY pue swd}sXsooq aunqesoqt] ayy JO MorAal e@ :sotoads oAtseAur onenbe Jo sjoedum o1wouocs oy], “TA ‘ZOIMOWEpy pue yy ‘oopreN, TS ‘surjymeyH pue ‘py ‘uosdmoyy “wy Jorsopung “q' “enes “wd ‘uossuog “y “SeyuwyH “gq ‘Bioey “JY sory “AH ‘uosourq “yf “Hoy[Ng “J ‘Typsoryporg “Pf ‘uosIspuy “TJ Iplomry “d “Bsoqy “W ‘HeIqqy “S'd “BIPYosoyy “CD ‘uosiajag puke “Vy ‘Za[BZUODH “PID “Uos|ayyIY TC AgNO Wd PAL pur “Ty Ike "f ‘seAeuog pur “5 ‘soyuoyy “g “zadoq-upeyy ‘f ‘seAeuog pure “> ‘sayuoyy “gq ‘zadoJ-ulpeyy Yy yeus Suey ‘DP WOATOY pue “CN “pos Por eR L syooug, *], SHY “fp ON|eE “H aupunsepy “DH soupy “JT ‘zapuvuio,y puv J's ‘aUo}g “P'S “][PAO7T] SYOH.LNV SOMMOUOIT AIANOSIY puv ]OIUAWUUOMAUT SCC SoAJosod JO Jaquinu pue ozis [eunjdo oy pue “son[BA O1WOUOSS dANjBIO1 ‘suoIe[ndod so1oads yo uoNquysip jenedg “WL Tsou! A pue “CW ‘shod OP8-O18 LOPI-L8e1 O9ST-LISI OPP-Sob S007 S007 S007 S007 PESO Pesenssy cs TOA p anssy TOA BULIIUIBUT [DISVOD suoyvoyddy poa1sojoaq SOIMOUOIT [PUA MUOMAUT yooqpuy Hy SIMMOUOIT AIANOSAY [OIUAULUOMAUA Solmouoda )P91BO]OIT uorjoajoid [B}seOd JOF spueWop [eBIOOS pue SIWOUOD SOAT}OALgO o1WOUODA puek Jeorsojo1qg YIM osn purer :advospuey Surom ev ur satoads Surasosuo| AjISIOAIPOTg JO sormouoody ay] 67 JoideyD {AssoAIporg jnoqge aio AT[ea1 OM OC] UOBOIFT}I9O 9ajJOO apeys pue “pyard “AysIOAIporg “Vy ‘Jop ua uaa, pur ‘g ‘Mjoziepy “gq ‘pWO[Og VW PIAHIEIS pue “XooTTA\ ‘dd PuApIO A ‘S ‘Hopuelye A TM S9eTTB ‘f 9Jenag pue “MM suyor “yy “PlozeM “WW “O]SYoo “MY “youqT aq ‘3 cu ‘O'D ‘Ape pue “yyy yowny, qng pure “> ‘umoig “yf “yoomary, OTIPPSEL BUY] pue ‘sg ‘Aysejog “q “ueUNyIy, 87 VY CuOse A pue ‘ff “Zormoyorys “WY z ra >: ySBOIAIOg UWa}SASOO — “PONG “A “BLY “Y'S “Wqwanjed “YSyo' YH *9Z10'T PS87I L00C STE TOA ELEY uvo0Q UO sso’'T AjyIsIaAtporg Jo sjouduy,, uo sjuaWWOD 0} osuodsay =D’ g'f “UOsyoRS “S'g “WodyeT] “D ‘oyfog “yf “AYN youuy “N quownrog “g'q Jsiqieg “gq “WHOA, “Y ‘uOsjeAy pur “pp “ZoImoyoRIS “WY 5 2 Zz o ‘ Ok “20¥]9S “A “BIBS “US TUNE “SYN MH “92107 O6L - LBL 9007 PIE PA aoualog SAOIAIDG W9}sKsooq UBD UO sso'T AjISIOAIPOIg JO sjoedul] Posh NGLEEU CVE ERI TIeT Cio) SON IE VLG yowmmg “N uounesg “gy Joqueg “gq “WOAY SW9Vd aLva WANTOA SYOHLNV 961-681 OP8-PE8 LOT-IST PTS-60S 9SI-Shl Sd9vd p-¢ Sonss] TIT TOA p-¢ sanss] OTT TOA t oss] OL TOA aALVa WIANNTOA 677 BuUIlJapoyY 1021801094 JUauluo.ilAUuy PUD sulajsdsooy ‘ainjjnoiay SJUAMUOALAUT pltp fo pousanor SOIMLOUOIT Jsaioy fo pouinor Aouog puv BIUALIY [PJUIUUOAAU Juauaspuryy pup aoualog Ayisiaalpolg fo ppousnor pouorpusajuy SONUOUOIT JOIMBOJOIT SOIMOUOIT [P9180J097 ADANOS SFY Woyynos Ur OZOT PUB OL6T USSMyeq uoleuoWsesy pue sso] jsasoy ozAjeue 0} sjopow jrordxe Ajjeneds 0002-0661 “sp4tq puejwney uvadosng Uo UOTBOLPISUdJUT [eIN]]NoLIse Jo sjoedun apim-jUsUT|UOD JO doUapPIAD JOYyVIN,| pueyenbewey ul asn pur] ajqeurejsns uo soanoadsiad Jeorsojoiq pue [eo1sojo1o0s Jo sisoyyuds W Heme] UI suaosaajno PIUOAIPY :sotoads SAISBAUL JOF SHOJFO [OIUOO WO sUOSSa] OIWOUODY a] Bs [BUOISAL B UO SouNseoUL [eyuoWUOLIAUS-Ude Are{UNJOA YIM OpIse-Jos JO UOTJeSI]gGO opim-ueadoing, ayy Suryury Aq Aorjod jeanjjnouse uowwoo uvodoing oy) Jo Surusess oy], sIsua]esuaq sdAp sunqjna paduini-oy1ym pososuepuo qussunuos so; suonevoydun uoneriosuoy BISy YyNog ul Ajeonto «jo uorenyea BIAT[OG ‘SOLB9N SOT Ul UOTOa}0Id poysiajzeM pue yeyqey psig Joy sjuowAed pury-uy :soorAsas [eyUOWUOSIAUD OM] BUTT[OS ([Ize1g) 9381 O[Ne OVS JsososUIEY JHuLy “yreg ALIS OGRIC] OP OLOPY JO UOHENTeA o1WOUOS| :pjiom suldojsasp ay} ul sos payoojoid Sunenyeas JOJ uONeNeA JUosuIUOS JO asn SY], O'V “UOIMAN “TN “TRH “WC “sawiooD “9 “BLU Na yoy fd “ura pourmog pue fT ‘plsyng fy “uosiopueg “yd “plevod eas ‘opyoy pue 'N ‘ddosipy “y JW “uBUJOH “gf “sursnoD ‘f ‘JOsseUNOY] puke “g “JosIey “yy “owing H 9yJaJd pure “H ‘9payoory ““H “siossog ‘DW WYg pue 'N “vurspuny, “Yy “wWeNeD “NY “JesEg “§ Uopun,y pue py ‘sesae, “P'N “yyinbsy Wid “Op W[essturg eprowyy pue “D euzy yoeqsioqy, “PPM WU “ZIHO “YU “BNOW ep aoueg “OD ‘sup SYOH.LAV SOIPTYS-9SLd O} Poy] SOPIUV EV GB LOE-18T SOb-LSP O1S-66b ETS-LOS TIL-689 “p-¢ sonssy ST TOA ¢§ onss] 6 TOA L007 LO0T [ anss] 9007 LIL TOA 9007 ASAMOH] ajqournjsng fo jouinor juauuUo.aug uvado.ing dood pun BUALIY [DJUAUUOMAUT (daa) Aono pojMaMUOAAUT uvadoing 1of ainjijsuy Sa1Mmouoay [0218010074 SOMOUOIT JOI30]OIA DOULA DAL UOlNPAAISUO jpasojoig Ay dpisoan peatoyst yy fo Jpusinor AoNOg asp) puoT ADINOS O€T BUeUR Ul S}ONPOIg ISO] PalfwaD :ATISIOAIPOTg JOF sjoyeyy 9QqIq JOATY ay} Suoje surefdpooy,y Jo uoNeIO}say ou} :spuepaA\ Jo uonenjeA olwoUucoY pue sAtoand Ylomowes Joye \\ uvadoing oy], sisA[eue [9pOUur JOJOAS ysouoy v :odoung] Ul UOIBAJASUOD jsaIOJ pasvosoUT Jo sjoedum o1WOUONy 1800/P007/VUA/T DANG "orssturw0 5 uvadoing ay} Joy Woda Jeuly “sooraios wojsksooa Jo sso] ay} 0} pa] sey sso] AYISIDAIPOIG aoyM sojdwexo Yq Suusumoog “Ajssoalporg Jo onjeA sasoos AIsAO00I Joy satoadg parosuepuy °¢*p Sursn Apnys-aseo V ‘sumeidoid AroAooar saroads pouajeosy) pue parosuepus woy sures] elensny jeoidoy ur puvy o}eAud uo uogseo posojsonbas -O1g JO} syiposo pue sjoyseu Surssowo jo Adojoo. pue AwouocW pue[uly ut uoBArasuoo AJISIDAIPOIG JSOOJ SUISBOIOUT JO 1OJOOS JSoJOJ OY} UO sjoeduT orWOUOSy OOIXayY “B10UOS Jo suteyd ouy ur AWAQonposd pue uoBAsasuoo puey (auniy19 wnjasiuuad) sseIsoyyng Bjewojeny Jo spuvlysty wi9}som oy) ul adeospue] Aqissoarp doo e yo Surdeys ayy :oziew Suiployy Spue]IOyJON SY} JOJ [00} poddns uorsioop ev Suipuayxq :Suruueyd Jeneds ut uontsodap uoLnu pue uonejuawsesy Jo sjs09 oy) SurzIPeWd}UT “A “RPA "V “Ipseyuyac] pur “f yousoAoyy “€ “Buaqiog pue -y “aaXasioyy “yy ‘onfey “[ UEHYY ‘d ‘yuug uo} pue "yy ‘uounyoy ‘oD ‘dedsuey pure ‘f jorpayioy ‘5 qunyy *V ‘ores |] yeeyy pue -y “uouruupyy] Wy ‘1enH pue -§[ ‘Moyleyy “Yy “Youwory-eurjoyy “q'y “uua[D “q ‘urndwey “Tq saysey “yy “suodT oy “upyuery ‘f “uBA UDA WW UMATTHH pur “PT “uosseay “Ap Jeeta A WYO ‘Uop ue yBIog “q ‘plemery “A a yurddy 99b-SSP PLL-19L SLI-691 80E-867 COP I-SShL OLE-S67 S00T 8007 8007 S007 LO0E p onss] 8ZT TOA c ans] fe JOA anss[ITE TOK anss] [Z JOA I Oss] 0S TOA I onss[ LZ TOA p-Z sonss] TL TOA UOIJMALBSUOD JDIISOJOIG JuaMadspUunyy /0IUAMUO.AU] auoolg AO aUuLinypy saigojoig snpuay sajduo ABOIOLG uoljyAsasuog unjayjng UONN[JO AULIOPY AydvisoaH payddy Buluun)d ungifQ) puv adnospunT OO aursnpy Lez ysewWUa(] UI UOT}OD]AS SAIOSOI [eUOIBOI Snsi9A JeuOeU Jo AoUdIOLJJO-1s00 oyen[eAd oO} sooud pueyuLey Sursy) pueyyyeoy ysiuec uo soroads pososuepuo Surjosyoid :Aouoyy Jo anye A, au0Z [e}seOD Ysipamg oy} ul uoudojaaap ayqeureysns 10F uoTeNTeA o1WOUOdy Bag UBAQqLIe dy} JO JUSWOSeUeUT a[qeuTEysNs Jo} SCIS Aq syusunMMIOS syqyy pue soroljod jo uoneurmexy adeospur] [einqjnoude youary v UI sjoosuT oNeNbe Jo AjISJOAIPOIG ay} O} UOI|NGLAUOD e ayeUT spuod ULE, Ajeqy Ul SIOATWeD 91] YIM sjor]FUOD uoTRAIasuO; Suypyoey, JO jsoD ayy pue Suruuejg uoNeAIosuoD oneUa}sXg slojeoIpul wr9}sAsooa outiew Jo juoudo]aAap 9] UO dAJodsiad YQ W Ope.udD UeIIzeIg Ul UONHBAIAsUOD AjISIOAIporq pue yUoUdoyaAap UeUINFY MN “optsdossoyy ur Apnjs v -asueyo JOAOO pur] pue osn pu] uUeqin Jo sjoedu JejUoWUONAUS SuTjOpoyy UOTJBAIASUOD FO sIJaUAq ay} pue olfloeg oy} ul Surysiy ourjsuo] jeusnpur seas ys1y Jo s}soo [e190s pue [eIN[No ‘orMOUOS ay] JO UOeSsaAUT UY :dUT] WOYOg oY], “AW puny] pue “yy “uosday “> “yoquey “N ‘adueng ‘d ‘dey, pure ‘pg ‘uosioyy, “gp ‘uasqooer “Ny ‘aduRS Vv “BAIOYNINOS pue “g “UOss|O “FY YWossy “] IstAbsapog “] ‘90y pue y “ysurg 'S “Hoqijosuy pue ‘d “Apeyy “f “vposonsiy “y ‘ourysarsD “vy ‘ouessny “| “tueylog pur “> “1uuipuoy “gq ‘Aemeudaly pur ‘['g ‘si9s0y CATAL suey “K Sulpjoy pure yy ‘souug “g Worneg YW °ZIPAO SIL9T-ZIL9T Sa9vd S007 aALVa Op onss] TOT TOA GANTIOA FDUNOS DAIOSOL JSAUOJUTEI UBOLIY Ue Je UOMBAIASUOD JO S]sOd Paadoxa AJISIOAIPOIG JO s}yousq o1WOUODy “TM ‘ZoIMOWepy pue “yy “OopleN ] anss] PIT TOA Snss]'Z9) [OA if anss] €T JOA 901-66 Sadovd WIANTOA JUBULUOALAUT pup Sula]SASOOT ‘ANNIE UOLINAAASUOD [DIISOJOLT AS24041 ajqournysng fo jpouinor SOMOUOIT [VIMBOJOIT aftpyin Jo suoisuauup ununpyy Aoyog auriopy ABO]007q paddy pun oispg AOUNOS CET seose [Bind sodoinq JO singny oy} aiojdxa Oo} quouidojaaap oueus9g Aorjod uoneasasuoo AyIsIoArporg uvadoing Jo ssaudAtjoajJO-}S09 dy] UO QpoIpe MOIAAI W {901NOsal a[qunyea v ButArosoid usyAr [nyaysem oq AYA\ poysioqe Ay jeueD eWeUeY 94] Ul SooIAJas Wo}sAsOod JO} sjayIeU SuNstXy solunos 1] SZ Joy soLeusos yuauIdojaacp ormouoss =pue jeonyod YM Sulapow Aysiaarporq = SuruTquiog (je) bIaq J9SIN JOUUT OY} UT spatq AJOeISTUE JO uorosjoid ay; Joy UOKRaJO joysePY :spsiq ALoyesSIU JO sormMoOUoDy sourddypiyd “org o8e1WOH POM OOSANN sJoou oyeyeqqny, Jo ase oY} :AJISIOAIPOIG OULIEUT JO soNfeA 1MIOUODS alnjded oO} suTsTURYDO;Y auljoap Jo sosnvo -adoing ut spuey jeursieyy Wf ‘soyyeg pue yy ‘uop uea Siog “ppy yoouysa Ay “yy Uoupsomyos pure * J ‘prozye Ay, “D ‘JEda0[Yog ouLAyVeD puke ‘Pf “URWISsOID “q ‘noyD “yy ‘uOsIopuy ““g “Yyorquoyne] “¢ “Jopueyype Ay “y ‘uoulieyy pue TW seszey “f “ulpy “Y ‘apewoypy “f “woogie, Hd ‘ues Suoynog pue “q ueruRyng “TU “opeqns ‘dq ‘pylins SHOH.LNV Annex 5 - Inventory of research organizations, programs and projects dealing with “Economics of Biodiversity Loss” As a contribution to the Scoping the Science project, an inventory was made of the main current research programs and organisations dealing with “economics of biodiversity loss” whereby 5 categories were distinguished: e 0 - Main networks/websites e 1 - Multi-lateral institutions, programs, treaties, etc e 2-Government supported initiatives e 3-NGO’s e 4- Universities / Research organizations & programs e 5- Business supported initiatives The inventory was carried out by using existing and well-known websites as a starting point (category 0 in below table). Most of these websites have a section with “links” and by following these in a systematic way it is possible to obtain a good overview of the main organizations and programs. Keywords used for the search were the link between “biodiversity” and “economics / economic loss / economic cost / economic valuation”. When going “deeper” into the various organisations and programs, many links are found to more specific subjects (e.g. programs, projects and organisations focusing on specific ecosystems (e.g. watershed-services, coral reefs, forests, etc), services or regions). Within the timeframe of this project it was not possible to explore these links in any detail but it would be very interesting and useful to carry out a full “network-analysis” as one of the future follow-up activities. Below, a summary is given of the main organisations and programs, organised by type of organisation. Per category a distinction is made between international and national organisations & programs. Within these categories, organisations are listed in alphabetical order for easy reference. This overview does not aim to be exhaustive, however in combination with the links given under the “Main Networks & Websites” (category 0) it should enable quick and easy access to the main organisations and programs on this topic. In the future, some resources could be directed to identifying relevant networks at national level. NOTE: In case you miss an (important) organization or program, please send an email to dolf.degroot@wur.nl] 0 - Main (long-term) networks / websites / databases (are also listed under respective country or organization for easy cross-reference) Economics World Bank Email: EnvEc@worldbank.org Ecosystem Service Database (ESD) http://esd.uvm/edu Ecosystem Services WWW.€Sa.Org/ ecoservices Ecosystem Services Database esd-worldbank.org/eei/ CSIRO Www.cse.csiro.au/ecoservices www.ecosystemvaluation.org Univ. Maryland www.epa.nsw.gov.au/envalue/ (start 1995) (EPA New South Wales) EVRI www.evri.ca Env. Canada & EPA-US (also (since 1997) World Bank & EC are involved) MILLENNIUM www.maweb.org ECOSYSTEM Natural Capital Project www.naturalcapital project.org Stanford University ASSESSMENT(2001-2005) Nature Valuation & Financing Network (NV&F) Ecosystem Services project (Australia) www.ecosystemservicesproject.org Ecosystem Valuation EnValue www.naturevaluation.org 1 — Multi-lateral Ints. / Org. / Programs / Treaties / etc. Name Website Additional information e Environmental Services and Z : Sustainable Use of Forests-program CIFOR / ICRAF www.cifor.cgiar.org e RUPES program (Rewarding Upland Poor for Environmental Services) Convention on_ Biological www.biodiversity.org Diversity - CBD or www.cbd.int o EGISS.GI ATES: e Env. Economics [program] ec.europa.eu/environment/enveco EXTERNE -— calculating the costs of externalities ec.europa.eu/research/ European Commission ec.europa.eu call for evidence on economics of biodiversity loss EURECA (European Ecosystem Assessment) European Environment Agency www.cea.curopa.eu www.eib.org Topic: Biodiversity (esp. irt health European Investment Bank loss); collaborate with IUCN e Financing Strategies for Sust. Forest Management (together with IUCN and CCAD) (www.fao.org/forestry/mecanismosfinancieros) MILLENNIUM ECOSYSTEM www.maweb.org ASSESSMENT (2001-2005) e Guidelines for Wetland Valuation (in Ramsar Convention www.ramsar.or: collab. with CBD) unstats.un.org/unsd/envaccoun e UN Committee of Experts on UN CEEA 5 : ; F ting/ceea Environmental Economic Accounting e Biodiversity portfolio (240 projects; Gustavo Fonseca) UNDP - GEF www.undp.org/gef UNEP Finance Initiative (www.unepfi.org) www.unep.ch UN Committee of Experts on Env. Economic Accounting (UNCEEA) WCMC (www.unep.wems.org) WWE UNESCO project about Financing Mechanisms for Protected UNESCO Www.unesco.org Areas (in Argentina) : www.vidasilvestre.org.ar UN-FCCC Impact CC on ecosystem services e Ecosystem Services Database — World Bank (esd-worldbank.org/eei/) e + environmental economics, environmental valuation, PES WORLD BANK www.worldbank.org 2 — Government supported initiatives [see also section 4: univ. and research programs] CSIRO Ecosystem Australia Services project [see also: main Networks] e Env. Canada & EPA-US (also World Bank & EC are involved) Additional information Www.cse.csiro.au/ecoservices www.evri.ca (since 1997) e established EEPSEA (Economy & Env. Program for SE Asia) in 1993 IDRC (Int. Dev. Research Centre) Canada www.idre.ca 235 supported by Royal Danish Min. of Foreign pANID AGES Germany GTZ MNP -_ Netherlands Environmental Assessment Agency Netherlands Ecosystem Services Project www.defra.goc.uk/environment Env. Economics Publ Atlas of Ecosystem Services Committee on Assessing www.nationalacademies.org and Valuing Ecosystem Services National Research Council 3 - NGO’s (organizations and programs) International Name Additional information of Association Environmental and www.aere.org Resource Economists e Started in 2002, last update 2006 - seems not very active anymore www.conservationfinance.org e Joint project of 12 org., among others: CI, IUCN, WWF, World Bank etc Conservation ‘cansenvaiGner mapping and valuing ecosystem International (C1) ME eee services ae e Economic analysis of benefits and deft org aa A : SEO TEIIE costs of biodiversity concervation RSITAS www.diversitas- DIVERS international.org East-West Centre Conservation Finance Alliance (CFA) e Ecoservices program (Charles Perrings) www.eastwestcenter.org European’ Centre for Nature Conservation (ECNC) rest Trends www. forest-trends.or. e see Katoomba Grou Fores 1) 5 > e development of payment i www.guianashield.or, e P Guyana Shield program mechanisms for ecosystem services 236 e Biodiversity for Sustainability program / Business & Biodiv. www.ecnc.nl IEEP (Inst. European Env. Policy) Wwww.ieep.eu e Env. Economics Program (EEP) e Support el. Newsletter on Payments for Watershed Services (www.flowsonline.net) www.lied.org e Funded by Env. Canada in 1990 IISD (Int. Inst. For Sust. Development) www.lisd.org) e Topics among others “Valuing Natural Capital” ISEE (Int. Society for Ecological Economics) www.ecoeco.org or www.ecologicaleconomics.org Support many activities irt Ecosystem Services and Economics: e Economics Division (since 1998) e Business and Biodiversity Program (since 2000) (cms.iucn.org/etc ..) e Global Economics and Env. Program (GEEP) > 2009 (cms.iucn.org/etc ..) e IUCN-Regional Env. Economics Program Asia ° WANI (PES Watershed services) (Wwww.iucn.org/themes/wani) e CEM-Ecosystem Services www.iucn.org e CEM-Ecosystem Approach www.katoombagroup.org [see also Forest Trends e Ecosystem Market Place (ecosystemmarketplace.com) Katoomba Group Restoring Natural Capital RRO TCA AHGeTOne (RNC) Alliance: i ae a a SANDEE (South Asian Network for Dev. and Env. Economics) www.sandeeonline.org Society for Ecological WWW. Ser.org, Restoration (SER). WCMC (see also UNEP) p.wemc.org e¢ mapping ecosystem services Wetlands International | wewetands.org | e Biorights program Worldl uiResources. | Inst e People and Ecosystems Program (WRI) (www.wri.org) www.wri.org/ecosystems/ecosyste m-services 237 Many activities eg: e (Centre for) Conservation Finance (www.worldwildlife.org/conservatio nfinance) e WWF Macro-economics Program Office www.panda.org/etc.. WWE www.worldwildlife.org e Meta-analysis of wetland values ... e Natural Capital Project (with Stanford Univ) e WWE UNESCO project about Financing Mechanisms for Protected Areas e §©6 Argentina: www. vidasilvestre.org.ar National NB: many national organizations were found but the list is rather incomplete and unbalanced and is therefore not included here (can be provided at request); an exception is made for the Ecological Society of America which made an interesting website with Toolkits on Ecosystem Services: www.esa.org/ecoservices 4 — Universities / Research programs (& assessments) International Website Additional information EVE concerted action www.landecon.cam.ac.uk/eve e EU funded project (2007-2009) e “Rationalising Biodiversity Conservation in Dynamic Ecosystems” RUBICODE www.rubicode.net National [STILL VERY INCOMPLETE] gene ee Australian Nat. University Economics and Environment Network Australia een.au.edu.au Australian Nat. University e a.o. host of global network of environmental economists 238 e PREM (Poverty Reduction and Netherlands - www.ivm.falw.vu.nl Env. Management) www.premonline.nl Netherlands | Tilburg Univ www.tilburguniversity.nl Et ee e SELS-program (www.ecosystemservices.nl): WUR (Wageningen research program on Ecosystem & Netherlands | Univ. & Research Wwww.wur.nl Landscape Services Centre) Wageningen International (www.wi.wur.nl): many projects CEEPA (Center for Env. South Africa | Pretoria Univ . -up.ac. Economics and Policy in Africa) (www.ceepa.co.za) Beijer Int. Inst. Environmental www.beijer.kwa.se Economics Env. Economics Unit Env. For Development Initiative (EfD) (www.efdinitiative.org) Sweden Gothenborg Univ. WWW.gu.se supported by SIDA Sweden Stockholm Univ ae Dept. Systems Ecology CSERGE Centre for Social & Economic Wwww.uea.ac.uk/cserge Research) UEA Wwww.greeneconomics- Green Economics Inst o org.uk e Center for Env. Management e Various projects, including MIMES; the Ecosystem Service Database (ESD) (esd.uvm/edu) Gund Inst. Ecological Economics-, UVM Snape eel e to be linked to ARIES-project (an interactive data base & DSS-tool), and the EcoValue Project (ecovalue.uvm.edu/evp) 239 e Ecosystem Valuation Maryland, Univ www.umd.edu www.ecosystemvaluation.org {__ Natural Capital Project (www.naturalcapitalproject.org) (G.Daily et al),together with The Nature Conservancy & WWF © Ini Stanford: Unis NatCap Network & Natural Capital Database InVEST — tool to model value of ES and Trade-offs Wyoming, Univ Dept. of Economics & | business.uwyo.edu/econfin Finance 5 - Business Supported Init. (Banks/ Consultants, etc ) Here only a few examples are listed but there are many more private organizations becoming increasingly involved in economics of biodiversity, and biodiversity loss “Building Biodiversity www.shell.com Business” Publication together with IUCN — 2008) Valuation Studies & Payment mechanisms TripleE www.tripleee.nl Economics for the Environment Consultancy www.Eftec.co.uk Ltd Exist since 1992, set up UKNEE (in 2004) Author Rudolf de Groot (Env. Systems Analysis Group, Wageningen Univ.) 240 Annex 6 - Key Internet Resources on Biodiversity and Ecosystem Services General: Biodiversity and Ecosystem Services Key institutions, groups and experts Center for Agri-Environmental Research, (University of Reading ) — The biodiversity and ecosystem services research group carries out rigorous research on the linkages between agricultural land-use, biodiversity, ecosystem function and service provision and how respond to global change and their value to man. http://www.reading.ac.uk/caer/theme_services.html Defenders of Wildlife — Defenders' Conservation Economics Program focuses on objective and transparent economic analysis of the benefits and costs of biodiversity conservation design of economic incentives for wildlife conservation Ecosystem Services. http://www.defenders.org/programs and_policy/science_and_economics/conservation_economics/inde x.php Defra’s Ecosystem Services Project (UK)- Defra research project on ecosystem services aims to establish the basis for an ecosystems approach and how it may be used to make effective assessments of the benefits that the natural environment provides. http://www.ecosystemservices.org.uk/index.htm Ecology and Society - An electronic, peer-reviewed, multi-disciplinary journal devoted to the rapid dissemination of current research on intergrative science for resilince and sustainability (formerly Conseration Ecology). http://www.ecologyandsociety.org/ Ecological Society of America and Union of Concerned Scientists. Ecosystem Services Toolkit — Tool for scientists to engage the public. Tool Kits have been completed on the following services: Pollination and Water Purification. Tool kit on Flood Damage Control is in development. http://www.esa.org/ecoservices/ Ecosystem Services Project - The Ecosystem Services Project was initiated by CSIRO Sustainable Ecosystems and The Myer Foundation. The Project is a collaborative natural resource management project studying the services people obtains from their environments, the economic and social values inherent in these services and the opportunities that can arise from considering these services more fully in land management policies and decisions. http://www.ecosystemservicesproject.org/html/aboutus/index.htm Ecosystem Services Management & Restoration - Online Database on Ecosystem Services: Linking Valuation and Financing of Ecosystem Services to Sustainable Management. This database on the website of the Nature Valuation & Financing (NV &F) network (www.naturevaluation.org) provides a list of ongoing and completed case studies, initiatives and projects from around the world that link the valuation and financing of ecosystem services to sustainable management. http://topshare.wur.nl/naturevaluation/73766 241 Millennium Ecosystem Assessment. - Designed by a partnership of UN agencies, international scientific organizations, and development agencies, this ongoing study assesses the capacity of ecosystems worldwide to provide goods and services that are important for human development. http://www.millenniumassessment.org/ Natural Capital Project. 2006. “Toolbox.” A joint venture among the Woods Institute for the Environment at Stanford University, the Nature Conservancy and the World Wildlife Fund is developing tools for modeling and mapping the delivery, distribution and economic value of ecosystem services and biodiversity. Online at: http://www.naturalcapitalproject.org/toolbox.htm| Global Biodiversity Outlook (GBO) - The GBO, is a periodic report on biological diversity, which provide a summary of the status of biological diversity and an analysis of the steps being taken by the global community to ensure that biodiversity is conserved and used sustainably, and that benefits arising from the use of genetic resources are shared equitably. http://www.cbd.int/gbo/ Global Environment Outlook (GEO) - The Global Environment Outlook (GEO) project is the implementation of UNEP’s mandate to keep the global environment under review. GEO is both a process and a series of reports, analyzing environmental change, causes, impacts, and policy responses. It provides information for decision-making, supports early warning and builds capacity at the global and sub-global levels. GEO is also a communication process that aims at raising awareness on environmental issues and providing options for action. http://unep.org/GEO/ Global Environment Outlook Year Book Series - The GEO Year Book series, produced annually by the United Nations Environment Programme in collaboration with many world environment experts keeps abreast of environmental issues as they unfold (i.e. an annual survey of the changing global environment). The Year Book includes global and regional overviews. It also highlights the most significant environmental developments in the year. http://www.unep.org/geo/yearbook/yb2008/ World Resources Institute — People and Ecosystems Programme. http://www.wri.org/ecosystems/ecosystem-services Reports, publications and journal articles Ecosystem Services: A Primer - Ecological Society of America (ESA) — This online article highlights the importance of natural ecosystems and the services they produce upon which humans are dependent. http://www.actionbioscience.org/environment/esa.html FAO. 2003. Biodiversity and the ecosystem approach in agriculture, forestry and fisheries Proceedings of the satellite event on the occasion of the Ninth Regular Session of the Commission on Genetic Resources for Food and Agriculture - Rome 12 - 13 October 2002. http://www.fao.org/docrep/005/y4586e/y4586e00.htm FAO.2007. The State of food and agriculture: Paying farmers for environmental services. Rome. Italy. http://www.fao.org/docrep/010/a1200e/a1200e00.htm Issues in Ecology, "Ecosystem Services: Benefits Supplied to Human Societies by Natural Ecosystems, No. 2, Spring, 1997, Ecological Society of America. Available on ESA's website at: http://www.esa.org/science_resources/issues/FileEnglish/issue2.pdf Mainka, S., McNeely, J and B, Jackson. 2005. Depend on Nature: Ecosystem Services Supporting Human Livelihoods. World Conservation Union. Available at: http://www.undp.org/pei/pdfs/IUCN- DependonNature.pdf Millennium Ecosystem Assessment, 2005, Living Beyond Our Means: Natural Assets and Human Well-being. Statement of the Board. http://www.millenniumassessment.org/documents/document.429.aspx.pdf Millennium Ecosystem Assessment, 2005. Ecosystems and Human Well-being: Wetlands and Water Synthesis. World Resources Institute, Washington, DC. http://www.millenniumassessment.org/documents/document.358.aspx.pdf Millennium Ecosystem Assessment, 2005. Ecosystems and Human Well-being: Synthesis. Island Press, Washington, DC. http://www.millenniumassessment.org/documents/document.356.aspx.pdf Millennium Ecosystem Assessment, 2005. Ecosystems and Human Well-being: Health Synthesis. Island Press, Washington, DC. http://www.millenniumassessment.org/documents/document.357.aspx.pdf Millennium Ecosystem Assessment, 2005. Ecosystems and Human Well-being: Desertification Synthesis. World Resources Institute, Washington, DC. http://www.millenniumassessment.org/documents/document.355.aspx.pdf Millennium Ecosystem Assessment, 2005. Ecosystems and Human Well-being: Opportunities and Challenges for Business and Industry. http://www.millenniumassessment.org/documents/document.353.aspx.pdf Naidoo, R and Ricketts, T.H. 2006. Mapping the Economic Costs and Benefits of Conservation PLoS Biology 4, 11. (2006). Available at: http://biology.plosjournals.org/perlserv/?request=getdocument&doi=10.1371/journal.pbio.0040360 RAND. Nature's Services: Ecosystems Are More Than Wildlife Habitat http://www.rand.org/scitech/stpi/ourfuture/NaturesServices/section|.html UNEP-WCMC. 2007. Biodiversity and Poverty Reduction. The importance of biodiversity for ecosystem services. Available at: http://www.unep- weme.org/latenews/Biodiversity%20and%20Poverty%20Reduction%20UNEP-WCMC. pdf World Bank, 2006. Where is the Wealth of Nations? Measuring Capital for the 21st Century. Washington, D.C.: The World Bank. http://siteresources. worldbank.org/INTEEI/214578- 11 10886258964/20748034/All.pdf World Resources Institute, 2005. World Resources 2005: The wealth of the poor — managing ecosystems to fight poverty. Washington DC: WRI. http://pdf.wri.org/wrr05_lores.pdf Giller, P., and G. O’Donovan, 2002. Biodiversity and Ecosystem Function: Do Species Matter? Proceedings of the Royal Irish Academy. 102B 3 129-139. http://www.ria.ie/cgi-bin/ria/papers/100066.pdf Crop pollination Key institutions, groups and experts CBD. Agricultural Biodiversity International Initiative for the Conservation and Sustainable Use of Pollinators. Established by the Fifth Conference of Parties to the Convention on Biological Diversity in 2000. Declared an urgent need to address the issue of worldwide decline of pollinator diversity? Available online at: http://www.cbd.int/programmes/areas/agro/pollinators.aspx Ecological Society of America (ESA). Pollination Fact Sheet: Pollination: an essential ecosystem service — Revealing secrets about the birds and the bees http://www.esa.org/ecoservices/comm/body.comm. fact.poll.html or http://www.esa.org/ecoservices/PollinationFactSheet.pdf European Pollinator Initiative (Hope Page) provides information on a wide range of activities which 243 helps to conserve and manage pollinators to enhance the servives they provide. http://www.europeanpollinatorinitiative.org/ IUCN. Task Force on Declinining Pollination of the Species Survival Commission of IUCN. Provides information on pollination as an ecosystem service in the conservation and sustainability of natural systems. http://www.uoguelph.ca/~iucn/ North American Pollinator Protection Campaign (Home Page) http://www.nappc.org/ The Arizona-Sonora Desert Museum, Forgotten Pollinators Campaign, Provides information on pollinators and links to key websites on this subject. www.desertmuseum.org/pollination/introduction.html USDA-ARS Bee Biology and Systematics Lab Research provide information based on various research projects on: the development and improvement of management systems for bee populations, biological studies of bees, plant-pollination systems, and bee biosystematics. www.LoganBeeLab.usu.edu/ Apimondia Journal. APIMONDIA (International Federation of Beekeepers' Associations) promote scientific, ecological, social and economic apicultural development in all countries and the cooperation of beekeepers’ associations, scientific bodies and of individuals involved in apiculture worldwide. Has various standing committees and publications on pollination. http://www.beekeeping.com/apimondia/index_us.htm Reports, publications and journal articles Alexandra-Maria Klein et al., 2007. Importance of pollinators in changing landscapes for world crops. Proc. R. Soc. B (2007) 274, 303-313. Available online at: http://www.environnement.ens.fr/perso/claessen/e3/tpe/kleinetal2006.pdf C. Kremen, N. Williams, M.A. Aizen, B. Gemmill-Herren, G. LeBuhn, R. Minckley, L. Packer, S. Potts, T. Roulston, I. Steffan-Dewenter, D. Vazquez, R. Winfree, L. Adams, E. Crone, S. Greenleaf, T. Keitt, A.M. Klein, J. Regetz, T. Ricketts. Pollination and Other Ecosystem Services Produced by Mobile Organisms: A Conceptual Framework for the Effects of Land Use Change. Ecology Letters 10, 4: 299-314. (2007) Available at: http://www.blackwell-synergy.com/doi/abs/10.1111/j.1461- 0248.2007.01018.x Daily, G.C, Ehrlich, P.R, and C.H. Sekercioglu. 2004. Ecosystem Consequences of Bird Declines. Proceedings of the National Academy of Sciences, USA 101, 52: 18042-18047. http://www.pnas.org/cgi/reprint/101/52/18042 Eardly, C .Pollinator —A conservation priority. Science on Africa Magazine. Available online at http://www.scienceinafrica.co.za/pollinator.htm Goulson, D. 2003. Conserving wild bees for crop pollination. Food, Agriculture & Environment Vol.1 (1): 142-144. Available online at: http://www.sbes.stir.ac.uk/people/goulson/intjourfoodagricenvir2003.pdf Kearns, C.A., and D. Inouye. 1997. “Pollinators, Flowering Plants and Conservation Biology,” BioScience 47: 297-397. Available at: http://www.jstor.org/pss/1313191 [Accessed 25 April 2008] Kremen, C., Williams, N. M , and R. W. Thorp. 2002. Crop pollination from native bees at risk from agricultural intensification. PNAS. December 24, 2002. Vol. 99, no. 26http://www.pnas.org/cgi/reprint/2624 13599v1 RAND. Pollination Services: No Food Without Them http://www.rand.org/scitech/stpi/ourfuture/NaturesServices/secl_pollinators.html Ricketts, T.H 2004. Tropical Forest Fragments Enhance Pollinator Activity in Nearby Coffee Crops. Conservation Biology 18, 5: 1262-1271. (2004). http://www.blackwell- synergy.com/doi/abs/10.1111/j.1523-1739.2004.00227.x APIMONDIA Standing Commission on Pollination and Bee Flora - 244 http://www.beekeeping.com/apimondia/index_us.htm Foraging Behaviour Of Honeybee On Parental Lines Of Hybrid Cauliflower Pusa Selvakumar. http://www.apimondia.org/apiacta/slovenia/en/selvakumar.pdf The Variability Of Yield Structure Of Black Currant Cultivars (Ribes nigrum L) In Different Pollination Conditions Denisow http://www.apimondia.org/apiacta/slovenia/en/denisow.pdf Induction Feeding Of Honey-Bees To Improve Actinidia deliciosa Pollination Gardi. http://www.apimondia.org/apiacta/slovenia/en/gardi.pdf Honeybee Pollination In Sunflower Hybrid Seed Production Yadav. http://www.apimondia.org/apiacta/slovenia/en/yadav.pdf Reforestation With Major Bee Food Trees In E] Salvador Sandker http://www.apimondia.org/apiacta/slovenia/en/sandker.pdf Genetic diversity of crops and livestock Key institutions, groups and experts Biodiversity International (Home Page)- Bioversity International is the world's largest international research organization dedicated solely to the conservation and use of agricultural biodiversity. http://www. bioversityinternational.org/. Carries research projects and publish extensively on genetic diversity of crops and livestock. http://www. bioversityinternational.org/Themes/Agricultural_ Ecosystems/index.asp UK Agricultural Biodiversity Coalition (UKabc) Home Page: http://www.ukabe.org/ukabe3.htm The EU Biodiversity in Development Project www.wemce.org.uk/biodev CGIAR agricultural biodiversity research centre Www.cgiar.org/ipgri Open Directory Project links on Agricultural Biodiversity www.dmoz.org/science/environment/biodiversity/agricultural Reports, publications and journal articles Daily, G.C, Ehrlich, P.R, Michener, C. and T.H. Ricketts. 2004. Economic Value of Tropical Forest to Coffee Production. Proceedings of the National Academy of Sciences, USA 101, 34: 12579-12582. Available at: http://www.pnas.org/cgi/reprint/101/34/12579 Cromwell, E., Cooper, D and Mulvany, P. Agriculture, biodiversity and livelihoods: issues and entry points for development agencies. http://www.ukabe.org/odi_agbiod.pdf FAO. 1997. The state of the world’s plant genetic resources for food and agriculture. Rome: Italy. Available at: http://www. fao.org/WAICENT/FAOINFO/AGRICULT/AGP/AGPS/Perfa/pdf/swrfull.pdf ITDG. Agricultural biodiversity: farmers sustaining the web of life. Farmer’s World Network Briefing. http://www. practicalaction.org/docs/advocacy/fwn_bio-div_briefing.pdf Sustaining Agricultural Biodiversity and Agro-ecosystem Functions: Opportunities, incentives and approaches for the conservation and sustainable use of agricultural biodiversity in agro-ecosystems and production systems. Report of the FAO/CBD Agricultural Biodiversity Workshop, 2-4 December 1998, Rome. http://www. fao.org/WAICENT/FAOINFO/SUSTDEV/EPdirect/EPre0063.htm Jn Situ Agricultural Biodiversity Conservation Project A research project of the Intermediate 245 Technology Development Group (ITDG) and the Overseas Development Institute, UK (ODI). http://www.ukabe.org/abe.htm Developing Diversity: European NGOs' PGRFA activities Illustrated keynote paper presented to the 1998 European PGRFA Symposium, Braunschweig, 30th June 1998. By Patrick Mulvany, Intermediate Technology, ITDG. http://www.ukabc.org/bschweigNGO.htm#p Breeds of Livestock resource presented by the Department of Animal Science at Oklahoma State University - an educational and informational resource on breeds of livestock throughout the world. http://www.ansi.okstate.edu/breeds/ Heifer Project International Information on livestock projects. http://www.heifer.org/ Marine Fisheries Key institutions, groups and experts Communication Partnership for Science and the Sea: Marine Ecosystem Services - COMPASS is a collaborative effort to advance marine conservation science and communicate scientific knowledge to policymakers, the public, and the media. http://www.compassonline.org/marinescience/ecosystem.asp http://www.compassonline.org/ FAO. Fisheries and Aquaculture Department ( Hope Page) —Inland Aquatic Ecosystems and Coastalk and Marine Information http://www.fao.org/fishery/ The FAO Fisheries and Aquaculture Department collects, analyzes and disseminates information on the sector operations (catch, production, value, prices, fleets, farming systems, employment). It also develops methodology, assesses and monitors the state of wild resources and elaborates resources management advice. http://www. fao.org/fishery/about World Resources Institute (WRI) 2006. The Value of coastal ecosystems http://earthtrends.wri.org/updates/node/118 Reports, publications and journal articles Duffy, J Emmett (Lead Author); Walker Smith (Topic Editor). 2006. "Marine ecosystem services." In: Encyclopedia of Earth. Eds. Cutler J. Cleveland (Washington, D.C.: Environmental Information Coalition, National Council for Science and the Environment). [First published in the Encyclopaedia of Earth October 10, 2006; Last revised October 16, 2006; Retrieved April 24, 2008]. http://www.eoearth.org/article/Marine_ecosystem_services Boris Worm, Edward B. Barbier, Nicola Beaumont, J. Emmett Duffy, Carl Folke, Benjamin S. Halpern, Jeremy B. C. Jackson, Heike K. Lotze, Fiorenza Micheli, Stephen R. Palumbi, Enric Sala, Kimberley A. Selkoe, John J. Stachowicz, and Reg Watson. Impacts of Biodiversity Loss on Ocean Ecosystem Services. Science 314, 5800: 787-790. (2006). http://www.sciencemag.org/cgi/content/abstract/3 14/5800/787 Issues in Ecology, " The Role of Nearshore Ecosystems as Fish and Shellfish Nurseries, No. 11, Spring, 2003, Ecological Society of America. Available on ESA's website at: http://www.esa.org/science resources/issues/FileEnglish/issue 1 1.pdf Timber from natural forests Key institutions, groups and experts The FAO Forestry Web site provides literally thousands of pages of information, access to all of FAOs 246 forest-related databases, detailed country profiles and links to documents on all aspects of forestry including new sites on forest fire, national forest programmes and forest reproductive material, among others. http://www. fao.org/forestry/en/ Information on specific forest products and services can be found at: 1. Wood energy http://www.fao.org/forestry/site/energy/en/ Pulp, paper and wood industries http://www.fao.org/forestry/site/harvesting/en/ 3. Trade in forest products and services http://www. fao.org/forestry/site/trade/en/ 4. Non-wood forest products http://www.fao.org/forestry/site/nwfp/en/ N Nature-related outdoor activities Reports, publications and journal articles e RSPB. 2007. Wellbeing through Wildlife in the EU. http://www.birdlife.org/eu/pdfs/Wellbeing EU_final_ version 2mb.pdf e Pretty J, Griffin M, Peacock J, Hine R, Sellens M and South N. 2005. A countryside for health and wellbeing: The physical and mental health benefits of green exercise. University of Essex, Colchester. Available online at: http://www.countrysiderecreation.org.uk/pdf/CRN%20exec%20summary.pdf http://195.92.230.85/Images/Hine_tcm2-3003 |.pdf e Bird W. 2004. Natural Fit: Can Green Space and Biodiversity Increase Levels of Physical Activity? RSPB. The RSPB. Available at: http://www.rspb.org.uk/Images/natural_ fit full version _tem9-133055.pdf e Collins, S. 2006. The Makuleke model for good governance and fair benefit sharing Steve Collins in IUCN. Policy Matters http://www.iucn.org/themes/ceesp/Publications/newsletter/PM 14-Section%201V.pdf e Health Council of the Netherlands. 2004. The influence of nature on social, physical and psychological wellbeing. Part 1: review of current knowledge. Report to the Minister of Agriculture, Nature and Food. http://www.rmno.ni/files_content/Nature%20and%20Health.pdf e UNEP/ CMS Secretariat (2006). Wildlife watching and tourism. Bonn. Available at: http://www.cms.int/publications/pdf/wildlifewatching text.pdf Soil Quality Key institutions, groups and experts e FAO. Soil Biodiversity Portal http://www. fao.org/ag/AGL/agll/soilbiod/index_en.stm Reports, publications and journal articles e Coleman, D.C and Whitman, W.B. 2004. Linking species richness, biodiversity and ecosystem function in soil systems. http://www. fao.org/ag/AGL/agll/soilbiod/promotxt.stmInterational symposium on impacts of soil biodiversity on biogeochemical processes in ecosystems, Taipei, Taiwan, 2004 http://www.umanitoba.ca/faculties/afs/soil_science/MSSS/Ecology/Graduate/Coleman%20and%20Whi tman%202005.pdf Biological control of crop pests and diseases Key institutions, groups and experts 247 HRDA. Natural Pest and Disease Control —Carries out research on biological control of pest, deseases and weeds. http://www.gardenorganic.org.uk/research/ir_pdw_man.php Natural hazard regulation Key institutions, groups and experts Dartmouth Flood Observatory keeps a Global Active Archive of Large Flood Events (1985-present) using information derived from a wide variety of news, governmental, instrumental, and remote sensing sources. International landslide centre is maintaining a worldwide landslide fatality database. The Hawai'i Solar Observatory has data on tropical storm paths worldwide stretching back to 1995, and a facility to calculate probability of strike given the co-ordinates of the location. Reliefweb records past natural disasters worldwide, including the extent of destruction and human suffering (but good maps of the area affected are not always available). UNEP-WCMC has compiled a World Atlas of Seagrasses (Green & Short 2003), and a World Atlas of Coral Reefs (Spalding et al. 2001). A World Mangrove Atlas (Spalding et al. 1997) is currently being updated. Global Lakes and Wetlands Database developed through a partnership between WWF and the University of Kassel in Germany (Lehner and D6ll 2004). Reports, publications and journal articles Bruijnzeel, L., van Dijk, A. I., van Noordwijk, M., Chappell, N. A., & Schellekens, J. 2007. Tropical deforestation, people and flooding. CIFOR. [online] http://www.worldagroforestry.org/water/downloads/bea_bruijnzeel.pdf Danielsen, F., M.K. Sorensen, M.F. Olwig, V. Selvam, F. Parish, N.D. Burgess, T. Hiraishi, V.M. Karunagaran, M.S. Rasmussen, L.B. Hansen, A. Quarto, and N. Suryadiputra. 2005. “The Asian Tsunami: A Protective Role for Coastal Vegetation.” Science. 310(5748): 643. Online at: http://www.sciencemag.org/cgi/content/abstract/3 10/5748/643 FAO, & CIFOR. 2005. Forests and floods. Drowning in fiction or thriving on facts? Bogor Barat, Indonesia: Center for International Forestry Research and Food and Agriculture Organization of the United Nations. Retrieved April 14, 2008, from http://www.fao.org/docrep/008/ae929e/ae929e00.htm MA. 2005. Regulation of Natural Hazards: Floods and Fires. http://www.millenniumassessment.org/documents/document.285.aspx.pdf NOAA. 2008. NOAA Pacific Tsunami Warning Center. http://www.prh.noaa.gov/pr/ptwe/ Medicinal Plant Species 248 Key institutions, groups and experts e Biodiversity and Human Health: http://www.ecology.org/biod/ e CIFOR Livelihood Briefs: Available at: http://www.cifor.cgiar.org/Publications/Briefs/Livelihoods/ e CIFOR Forest and Health Initiative - http://www.cifor.cgiar.org/Research/Livelihoods/MainActivities/F orestHealth/introduction.htm Reports, publications and journal articles e Chivian E. et al., (2004) Biodiversity: Its importance to human health. The Center for Health and the Global Environment Harvard Medical School. Available online at: http://chge.med.harvard.edu/publications/documents/Biodiversity_v2_screen.pdf e Colfer, C.J.P., Sheil, D and M. Kishi. Forests and Human Health: Assessing the Evidence Center for International Forestry Research (2006) Available at: http://www.cifor.cgiar.org/publications/pdf_files/OccPapers/OP-45.pdf Husselman, M. 2008. Beekeeping in Zambia. CIFOR Livelihood Brief 7. 4p. http://www.cifor.cgiar.org/publications/pdf_files/livebrief/livebrief0801.pdf Shackleton, S.; Kaschula, S.; Twine, W.; Hunter, L.; Holding-Anyonge, C.; Petheram, L. 2006. Forests as safety nets for mitigating the impacts of HIV/AIDS in southern Africa. CIFOR Livelihood Brief 4. 4p. http://www.cifor.cgiar.org/publications/pdf_files/livebrief/livebrief0604e.pdf Colfer, C.J.P.; Sheil, D.; Kishi, M.2006. Forests and human health: assessing the evidence. CIFOR Occasional Paper No. 45. Center for International Forestry Research (CIFOR), Bogor, Indonesia. 11 1p. Available: http://www.cifor.cgiar.org/publications/pdf_files/OccPapers/OP-45.pdf Colfer, C.J.P.; Sheil, D.; Kaimowitz, D.; Kishi, M.2006. Forests and human health in the tropics: some important connections. Unasylva 57(224): 3-10 online [URL] ftp://ftp.fao.org/docrep/fao/009/a0789e/a0789e02.pdf and http://www.fao.org/docrep/009/a0789e/a0789e00.htm Koppert, G.J.A, Dounias, E, Froment, A and Pasquet, P. 1993. Food consumption in the forest populations of the southern coastal area of Cameroon: YASSA — MVAE — BAKOLA. In: Hladik, C.M., Hladik, A., Linares, O.F. Pagezy, H., Semple, A. and Hadley, M. (1993). Tropical Forests, People and Food. Biocultural Interactions and Applications to Development. Man and the Biosphere Series, 13 (ParisE: UNESCO et CamforthE: The Parthenon Publishing Group), pp. 295-310. Available online at: http://www.cifor.cgiar.org/publications/pdf _files/research/forests_health/17.pdf Shanley, P and Luiz, L. 2003. The Impacts of Forest Degradation on Medicinal Plant Use and Implications for Health Care in Eastern Amazonia. Vol. 53 No. 6 BioScience 573. June 2003. Available online at: http://www.cifor.cgiar.org/publications/pdf_files/research/forests health/22.pdf Non use values Key institutions, groups and experts e Amazon Conservation Team - work in partnership with indigenous people in conserving biodiversity, health, and culture in tropical America. http://www.amazonteam.org/publications.html e Sacred Mountains Program is a program of the Mountain Institute's which mission it is to advance mountain cultures and preserve mountain environments. http://www.mountain.org/work/sacredmtns/index.cfm e The Theme on Indigenous and Local Communities, Equity, and Protected Areas (TILCEPA), was set up in 2000 by the World Commission on Protected Areas (WCPA) and the Commission on 249 Environmental, Economic, and Social Policy (CEESP) of the World Conservation Union (IUCN). It advocates, in all countries, the recognition of community conserved and managed areas that are significant from biodiversity point of view, and the development of management partnerships with the communities resident in or surrounding official PAs. http://www.iucn.org/themes/ceesp/Wk /TILCEPA/TILCEPA.htm The IUCN Commission on Environmental, Economic and Social Policy (CEESP), is an inter- disciplinary network of professionals whose mission is to act as a source of advice on the environmental, economic, social and cultural factors that affect natural resources and biological diversity and to provide guidance and support towards effective policies and practices in environmental conservation and sustainable development. http://www.iucn.org/themes/ceesp/index.html The Forum on religioin and Ecology is the largest international multireligious project of its kind. With its conferences, publications, and website it is engaged in exploring religious worldviews, texts, and ethics in order to broaden understanding of the complex nature of current environmental concerns. http://environment.harvard.edu/religion/main.html IUCN WCPA Task Force_on Cultural and Spiritual Values of Protected Areas (CSVPA) http://www.iucn.org/themes/wepa/theme/values/values.html World Water Day. 2006. Facts and figures about water religions and beliefs. http://www.worldwaterday.org/page/442 Freshwater Key institutions, groups and experts Ecological Society of America (ESA). Water purification fact sheet: Water purification an essential ecosystem services - Revealing secrets about natural water purification http://www.esa.org/ecoservices/comm/body.comm.fact.wate.html http://www.esa.org/ecoservices/WaterPurificationFactSheet.pdf Environment Canada: The Great Lakes Fact Sheet http://www.on.ec.gc.ca/wildlife/factsheets/fs_wetlands-e.html The hydrology module of the Natural Capital Project is being developed by Sue White at Cranfield University and Guillermo Mendoza at Stanford University. Reports, publications and journal articles Alcamo, J., Vuuren, D. V., Ringler, C., Cramer, W., Masui, T., Alder, J., & Schulze, K. 2005. Changes in nature’s balance sheet: model-based estimates of future worldwide ecosystem services. Ecology and Society 10: 19. [online] URL: http://www.ecologyandsociety.org/voll 0/iss2/art19/ Issues in Ecology, " Sustaining Healthy Freshwater Ecosystems. No. 10, Winter, 2003, Ecological Society of America. Available on ESA's website at http://www.esa.org/science_resources/issues/FileEnglish/issue 10.pdf RAND. Nature's Services: Ecosystems Are More Than Wildlife Habitat: Watershed - http://www.rand.org/scitech/stpi/ourfuture/NaturesServices/sec]_watershed.html Inland fisheries Key institutions, groups and experts Environment Canada: The Great Lakes Fact Sheet http://www.on.ec.gc.ca/wildlife/factsheets/fs_wetlands-e.html FAO. Fisheries and Aquaculture Department ( Hope Page) — nland Aquatic Ecosystems and Coastalk 250 and Marine Information: http://www. fao.org/fishery/ Reports, publications and journal articles e Issues in Ecology, " Effects of Aquaculture on World Fish Supplies. No. 8, Winter, 2001, Ecological Society of America. Available on ESA's website at: http://www.esa.org/science_resources/issues/FileEnglish/issue8.pdf Wild Meat Key institutions, groups and experts e Bushmeat Trade POSTnote, Parliamentary Office of Science and Technology. http://www.parliament.uk/documents/upload/POSTpn236.pdf e Conservation Science Group, Imperial College. http://www.iccs.org.uk/ e ODI Wild Meat, Livelihoods Security and Conservation in the Tropics. This research project focus on the human and social dimensions of hunting wild meat for consumptive use in tropical forests. http://www.odi-bushmeat.org/#home_research e UK Tropical Forest Forum Bushmeat Working Group. http://www.forestforum.org.uk/tradee.htm e WCS Hunting and Wildlife Trade Program. World Conservation Society Hunting and Wildlife Trade Programme. http://www.wes.org/international/huntingandwildlifetrade e ZSL Bushmeat and Forests Conservation Programme. http://www.zsl.org/field-conservation/bushmeat- and-forest/ Reports, publications and journal articles e DFID, Wildlife and Poverty Study, DFID Livestock and Wildlife Advisory Group London, 2002. http://www.dfid.gov.uk/pubs/files/wildlifepovertystudy.pdf e Bowen-Jones, E. What are the impacts of the bushmeat trade on biodiversity, and what entry points can the EU most effectively use to reduce these. European Association of Zoos and Aquaria (EAZA). Available online at: http://www.eaza.net/download/summebj.PDF Livestock Key institutions, groups and experts e Agriculture and Agri-food Canada: Biodiversity and Livestock Production http://www.agr.gc.ca/pfra/biodiversity/grazing_e.htm e The UK agricultural biodiversity coalition (UKabc) www.ukabc.org e The EU Biodiversity in Development Project. www.wemce.org.uk/biodev e CGIAR agricultural biodiversity research centre www.cgiar.org/ipgri e Open Directory Project links on Agricultural Biodiversity. www.dmoz.org/science/environment/biodiversity/agricultural Reports, publications and journal articles e DFID, Wildlife and Poverty Study, DFID Livestock and Wildlife Advisory Group London, 2002. http://www.dfid.gov.uk/pubs/files/wildlifepovertystudy.pdf Global Climate regulation 251 Key institutions, groups and experts e CIFOR. TroFCCA: Tropical Forests & Climate Change Adaptation Project. http://www.cifor.cgiar.org/trofcca/_ref/home/index.htm e The World Agroforestry Center (ICRAF). http://www.worldagroforestry.org/sea/Networks/RUPES/index.asp e IUCN Forests and Climate Change Initiative. http://cms.iucn.org/about/work/programmes/forest/fp_our_work/fp_our_work_thematic/fp_our_work_f cc/index.cfm e UNEP-WCMC. Biodiversity and Climate Change. http://www.unep-weme.org/Climate/ Authors Abisha Mapendembe & Matt Walpole (UNEP-WCMC, Cambridge, UK) D2, 4 ie